paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
astro-ph/0008262
1
0008
2000-08-17T05:49:07
Effects of dust extinction on optical spectroscopic properties for starburst galaxies in distant clusters
[ "astro-ph" ]
Recent observational studies on galaxies in distant clusters discovered a significant fraction of possible dusty starburst galaxies with the so-called `e(a)' spectra that are characterized by strong H$\delta$ absorption and relatively modest [OII] emission. We numerically investigate spectroscopic and photometric evolution of dusty starburst galaxies in order to clarify the origin of the e(a) spectra. We found that if a young starburst population is preferentially obscured by dust than an old one in a dusty starburst galaxy, the galaxy shows e(a) spectrum. It is therefore confirmed that the selective dust extinction, which is first suggested by Poggianti & Wu (2000) and means the strongest dust extinction for the youngest stellar population among stellar populations with different ages, is critically important to reproduce quantitatively the observed e(a) spectra for the first time in the present numerical study. We furthermore discuss what physical process is closely associated with this selective dust extinction in cluster environment.
astro-ph
astro-ph
Effects of dust extinction on optical spectroscopic properties for starburst galaxies in distant clusters Yasuhiro Shioya Astronomical Institute, Tohoku University, Sendai, 980-8578, Japan and Kenji Bekki Division of Theoretical Astrophysics, National Astronomical Observatory, Mitaka, Tokyo, 181-8588, Japan Received ; accepted 0 0 0 2 g u A 7 1 1 v 2 6 2 8 0 0 0 / h p - o r t s a : v i X r a -- 2 -- ABSTRACT Recent observational studies on galaxies in distant clusters discovered a significant fraction of possible dusty starburst galaxies with the so-called 'e(a)' spectra that are characterized by strong Hδ absorption and relatively modest [OII] emission. We numerically investigate spectroscopic and photometric evolution of dusty starburst galaxies in order to clarify the origin of the e(a) spectra. We found that if a young starburst population is preferentially obscured by dust than an old one in a dusty starburst galaxy, the galaxy shows e(a) spectrum. It is therefore confirmed that the selective dust extinction, which is first suggested by Poggianti & Wu (2000) and means the strongest dust extinction for the youngest stellar population among stellar populations with different ages, is critically important to reproduce quantitatively the observed e(a) spectra for the first time in the present numerical study. We furthermore discuss what physical process is closely associated with this selective dust extinction in cluster environment. Subject headings: galaxies: clusters -- galaxies: formation -- galaxies: ISM -- galaxies: infrared -- galaxies: interaction -- galaxies: structure -- 3 -- 1. Introduction Since Butcher & Oemler (1978) discovered a large fraction of blue galaxies in distant clusters of galaxies at the redshift z ≥ 0.2, the origin of the blue color and physical processes closely associated with the formation of the distant blue populations have been extensively discussed by many authors (Dressler & Gunn 1983, 1992; Lavery & Henry 1986; Couch et al. 1994, 1998; Abraham et al. 1996; Barger et al. 1996; Fisher et al. 1998; Morris et al. 1998; Balogh et al. 1999). In particular, Dressler & Gunn (1983, 1992) discovered galaxies with no detectable emission lines and very strong Balmer absorption ones (the so-called E+A galaxies) and concluded that some fraction of distant blue populations are poststarburst galaxies that truncated abruptly their active star formation. One of longstanding and remarkable problems concerning the evolution of galaxies in cluster environment is to understand clearly a physical relationship between these blue populations, whose number fractions rapidly increase with redshift, and passive populations such as elliptical galaxies and S0s. Several attempts have been made to clarify an evolutionary link among star-forming, poststarburst, and passively evolving galaxies observed in distant clusters of galaxies (Couch & Sharples 1987; Abraham et al. 1996; Barger et al. 1996; Morris et al. 1998; Balogh et al. 1999). It is however not so clear whether there is really an evolutionary link among these populations with various spectroscopic properties and what physical process can drive the evolution. Recent observational studies have found a significant population of possible dusty starburst galaxies in distant clusters (Poggianti et al. 1999: Owen et al. 1999; Smail et al. 1999), which provide a new clue to the evolutionary link among various spectral classes. Poggianti et al. (1999) suggested that galaxies with strong Balmer absorption and relatively modest [OII] emission, which are classified as "e(a)" galaxies by Dressler et al. (1999), are dusty starburst galaxies and furthermore discussed that distant clusters -- 4 -- have a significant fraction of these dusty starburst galaxies. Smail et al. (1999) found that spectral properties of 5 out of 10 galaxies detected by 1.4 GHz VLA radio observation are classified as poststarburst (a+k/k+a type in Dressler et al. 1999) and considered that the star formation is hidden by dust in them. This finding suggested that dust effects are remarkable even for galaxies that were previously identified as poststarburst ones by optical spectral properties. Although a growing number of observational results on e(a) galaxies with possible dusty starburst have been accumulated, there are only a few theoretical studies addressing the formation and evolution of these e(a) populations in distant clusters. The purpose of this Letter is to investigate the origin and the nature of e(a) population observed in distant clusters of galaxies by using a one-zone chemical and spectrophotometric evolution model. We particularly investigate whether dusty starburst galaxies can show both strong Hδ absorption and relatively modest [OII] emission observed in distant e(a) galaxies. We demonstrate that if a younger stellar population formed during a starburst is preferentially obscured by dust in a galaxy, spectral properties of the galaxy during starburst become very similar to those characteristics of e(a) populations. We refer to this behavior of dust extinction as a selective dust extinction, which is originally proposed by Poggianti and Wu (2000) and means that the effect of dust extinction is maximum for the youngest stellar populations among stellar populations with different ages. We furthermore suggest that this selective extinction can be achieved in interacting and merging galaxies in distant clusters. In the followings, the cosmological parameters H0 and q0 are set to be 65km s−1 Mpc−1 and 0.05 respectively, which means that the corresponding present age of the universe is 13.8 Gyr. -- 5 -- 2. Model We adopt a one-zone chemical and spectrophotometric evolution model of a disk galaxy with a starburst, and thereby investigate when and how a disk galaxy shows strong Hδ absorption line and relatively modest [OII] emission one characteristics of e(a) spectra observed in distant clusters of galaxies. Since more details of the adopted model are given in Shioya & Bekki (1998, 2000 in preparation), we only briefly describe the model in the present study. We follow the chemical evolution of galaxies by using the model described in Matteucci & Tornamb`e (1987) which includes metal-enrichment processes of type Ia and II supernovae (SNIa and SNII). We adopt the Salpeter initial mass function (IMF), φ(m) ∝ m−1.35, with upper mass limit Mup = 120M⊙ and lower mass limit Mlow = 0.1M⊙. We calculate photometric properties of galaxies as follows. The monochromatic flux of a galaxy with age T , Fλ(T ), is described as Fλ(T ) = Z T 0 FSSP,λ(Z, T − t)ψ(t)dt, (1) where FSSP,λ(Z, T − t) is a monochromatic flux of a single stellar population with age T − t and metallicity Z, and ψ(t) is time-dependent star formation rate described later. In the present study, we use the spectral library GISSEL96 which is the latest version of Bruzual & Charlot (1993). The star formation history of a disk galaxy is characterized by three epochs. The first is the epoch of galaxy formation (zform in redshift and the age of galaxy is 0 Gyr) at which a disk galaxy forms and begins to consume initial interstellar gas by star formation with the moderate rate. In the present study, zform is fixed at 4.5 and therefore the age of galaxies at z = 0 is 12 Gyr. The second is zsb at which starburst begins in the disk. In the present study, zsb is set at 0.4 and the age of galaxies at zsb, Tsb, is 7.64 Gyr. The third is zend at which star formation ceases and is defined as the epoch at which stellar mass fraction becomes 0.95 in our models. The age of galaxies at zend, Tend, is 8.56 Gyr. In the following, -- 6 -- we mainly use the age of galaxies to describe their evolution. Throughout the evolution of disk galaxies, the star formation rate is assumed to be proportional to gas mass fraction (fg) of galaxies; where k is a parameter which controls the star formation rate. This parameter k is given as ψ(t) = kfg, (2) follows: k = kdisk for 0 ≤ T < Tsb, ksb 0 for Tsb ≤ T < Tend, for Tend ≤ T. (3)   In the following, we refer the first, second, and third phases as the prestarburst, the starburst, and the poststarburst, respectively. In the present study, we set the value of kdisk as 0.225 Gyr−1, which corresponds to a plausible star formation rate for Sb disks (e.g., Arimoto, Yoshii, & Takahara 1992). ksb is considered to be 10 times larger than kdisk (2.25 Gyr−1), which is consistent with observational results on starburst galaxies (e.g., Planesas et al. 1997). Figure 1 shows the star formation history of our model. EDITOR: PLACE FIGURE 1 HERE. We try to understand the role of the selective dust extinction in the formation of e(a) galaxies by comparing the following two models with each other; a model with no dust extinction (from now on referred to as the ND model) and the selective dust extinction model (the SD model). In the ND model, dust effects on photometric and spectroscopic properties are completely neglected; i.e., AV = 0 at all times. In the SD model, we assume that during starburst the value of AV depends on the age of stellar population; i.e., AV = AV,0 exp{(T − t)/τ } for Tsb < T < Tend, (4) where AV,0 and τ are parameters controlling the degree of extinction. We here set AV,0 = 5 mag and τ = 1.0 × 106 yr. -- 7 -- Based on the monochromatic flux derived in the equation (1) and the value of AV from the equation (4) for each age of a galaxy T , we calculated the SED of a galaxy corrected by dust extinction using the extinction law derived by Cardelli et al. (1989) and adopting the so-called screen model. For deriving the fluxes for various gaseous emission lines (Hδ and [OII]) in dusty starburst galaxies, we first calculate the number of Lyman continuum photons, NLy, by using the SED that are not modified by dust extinction. If all the Lyman continuum photons are used for ionizing the surrounding gas, the luminosity of Hβ is calculated according to the following formula; L(Hβ)(erg s−1) = 4.76 × 10−13NLy(s−1) (5) (Leitherer & Heckman 1995). To calculate luminosities of other emission lines, e.g., [OII] and Hδ, we use the relative luminosity to Hβ luminosity tabulated in PEGASE (Fioc & Rocca-Volmerange 1997) which is calculated for the set of electron temperature of 10000 K and electron density of 1 cm−3. Thus the SED derived in the present study consists of stellar continuum and gaseous emission. EDITOR: PLACE FIGURE 2 HERE. EDITOR: PLACE FIGURE 3 HERE. 3. Results Figure 2 shows the time evolution of EW(Hδ) and EW([OII]) for the ND model and the SD one. First, we describe the evolution for the ND model. The equivalent width of [OII] decreases gradually before the starburst begins (i.e., at T < 7.64 Gyr). At T = Tsb, -- 8 -- the equivalent width of [OII] increases rapidly following the rapid increase of star formation rate due to starburst and then decreases gradually with the decreasing star formation rate. During the starburst, the equivalent width of [OII] is larger than 40 A. After star formation ceases at T = Tend and massive stars formed during starburst die out, the equivalent width of [OII] rapidly decreases and becomes smaller than 5 A. The strength of equivalent width of Hδ evolves as follows. During the prestarburst evolution phase, the Hδ line is observed as an absorption one and its equivalent width reaches the local maxima at around 3 Gyr (z ∼ 1.5). The strength of EW(Hδ) gradually decreases, and becomes about 3 A just before starburst begins (T = 7.64 Gyr, z = 0.4). When starburst begins, the Hδ line changes from an absorption line to an emission one and its equivalent width reaches −5 A. A few × 108 yr after the starburst begins, Hδ line changes into absorption line again. Since the flux of Hδ emission line becomes negligibly small, the equivalent width of Hδ reaches 8 A when starburst ceases. As a result of passive evolution of galaxies after starburst, the equivalent width of Hδ decreases gradually: it becomes smaller than 7 A at 8.8 Gyr and 3 A at 11 Gyr. Next we describe the evolution of the SD model. Since we assume that the selective extinction affects spectroscopic properties of galaxies only in starburst phase, there is no difference in the evolution of EW([OII]) and EW(Hδ) between the SD model and the ND one in the prestarburst phase and the poststarburst one. We therefore pay our attention to the evolution of EW([OII]) and EW(Hδ) during starburst. In the SD model, the equivalent width of [OII] does not increase (but increases in the ND model) and is below 40 A during starburst. The equivalent width of Hδ becomes rapidly small just after starburst begins, although the Hδ line can be still observed as an absorption line. A few ×108 years after the starburst begins, the equivalent width of Hδ becomes larger than 4 A. The reason why the equivalent width of Hδ of the SD model is always larger than that of the ND model is that the flux of Hδ emission line of the SD model is smaller than that of ND model. -- 9 -- Figure 3 shows the evolution of galaxies on EW([OII])-EW(Hδ) plane for each model. The criteria of spectroscopic classification determined by Dressler et al. (1999) are also superimposed on it. The ND model is located within the e(b) region during starburst and is very rapidly shifted to the a+k one just when starburst is ended. Although, the locus of the ND model passes through the e(a) region, the crossing time is 4 × 106 years which is about 0.5 % of the duration of the starburst. On the other hand, the SD model can successfully show the spectroscopic properties of e(a) galaxies, namely, strong Hδ absorption line [EW (Hδ) > 4 A] and relatively modest [OII] emission line [EW([OII]) > 5 A], during the most of the starburst phase (0.7 Gyr). This result accordingly suggests that a selective extinction plays a vital role in the formation of galaxies with e(a) spectra. These results thus confirm the early suggestion by Poggianti & Wu (2000) that if the youngest stellar population is the most heavily obscured by dust in a dusty starburst galaxy, the e(a) spectra can be achieved in the galaxy. As is described above, we have confirmed that the SD model shows the weaker [OII] emission and stronger Hδ absorption compared with the ND model. The physical reason for the successful reproduction of the SD model can be understood in terms of the difference in the influence of dust extinction between continuum and emission lines. The emission lines come from the ionized gas around youngest stellar populations which are most heavily obscured by dust. On the other hand, most of the continuum come from the stellar populations that are older and less obscured than younger ones dominating the ionizing photons. The flux of emission lines consequently is more greatly affected by dust extinction than that of stellar continuum. Thus the SD model can show the smaller equivalent width of emission lines without changing so greatly the flux of continuum. -- 10 -- 4. Discussion and conclusions Poggianti & Wu (2000) first discussed that if the dust extinction for a stellar population decreases with the stellar age, e(a) spectra can be obtained for a starburst galaxy. They furthermore suggested that if the location and thickness of dust patches depend on the age of the embedded stellar populations in a starburst galaxy, the effects of the above selective dust extinction become remarkable for the galaxy. Then, when and how the proposed difference in the degree of dust extinction between stellar populations with different ages is possible during galactic evolution in distant clusters? We here suggest that the difference of dust extinction between the central region and the outer one in a galaxy is a main cause for the above age-dependent extinction. To be more specific, since younger stellar populations formed by secondary starburst in the central region with a larger amount of dusty interstellar gas are more heavily obscured by dust than the outer old components, the age-dependent extinction can be achieved. This radial dependence of dust extinction is considerably reasonable and realistic, considering that secondary dusty starburst occurs preferentially in the central part of a galaxy owing to efficient inward gas transfer driven by non-axisymmetric structure (such as stellar bars) and galaxy interaction and merging. Numerical simulations on gaseous and stellar distribution in merging disk galaxies with dusty starburst furthermore have demonstrated that the central young stellar component formed by nuclear starburst is more heavily obscured by dusty gas than the outer old component initially located in merger progenitor disks (Bekki, Shioya, & Tanaka 1999). Thus we strongly suggest that galaxies with the central dusty starburst are more likely to show the selective dust extinction. Future observational studies on the radial dependence of photometric and spectroscopic properties for e(a) galaxies, which can reveal the detailed distribution of young dusty population and that of old one with less extinction, will assess the validity of the above idea for the origin of the selective dust extinction (i.e., the age-dependent extinction). -- 11 -- Then what physical process is closely associated with the radial dependence of dust extinction suggested above and thus with the formation of e(a) galaxies in distant clusters ? We here suggest that minor and unequal-mass galaxy merging, which is different from major merging in that it can leave disk systems even after merging, is one of possible candidates that can reproduce fundamental properties of e(a) galaxies. Mihos & Hernquist (1994) has demonstrated that minor merging can excite non-axisymmetric structure in a late-type disk, trigger nuclear starburst, and transform the disk into an early-type disk such as Sa and S0. Unequal-mass merging between two disks is furthermore demonstrated to create a disk with the structure and kinematics strikingly similar to those of S0s (Barnes 1994), and furthermore Bekki (1998) found that such merging can trigger nuclear starburst, form a very gas-poor disk, and create a remarkable bulge characteristic of typical S0s. These merging therefore not only contribute to the formation of dusty nuclear starburst and thus to the radial dependence of dust extinction but also provide an evolutionary link between late-type disks, e(a) galaxies, and early-type disks (S0) with the number fraction observed to decrease with redshift (e.g., Dressler et al. 1996). An early-type disk galaxy formed by minor and unequal-mass merging experiences the post-starburst phase with k+a/a+k spectra after starburst and shows a remarkable bulge with young stellar populations and thus with bluer colors (Bekki 1998). Accordingly one of observational tests for assessing the relative importance of minor and unequal-mass merging in the fraction of e(a) galaxies with the selective dust extinction is to investigate the number fraction of disk galaxies with k+a/a+k spectra and prominent blue bulges in distant clusters. We are grateful to the anonymous referee for valuable comments, which contribute to improve the present paper. Y.S. thanks to the Japan Society for Promotion of Science (JSPS) Research Fellowships for Young Scientist. -- 12 -- REFERENCES Abraham, R. G., Smecker-Hane, T. A., Hutchings, J. B., Carlberg, R. G., Yee, H. K. C., Ellingson, E., Morris, S., Oke, J. B., Rigler, M. 1996, ApJ, 471, 694 Arimoto, N., Yoshii, Y., & Takahara, F. 1992, A&A, 253, 21 Balogh, M. L., Morris, S. L., Yee, H. K. C., Carlberg, R. G., & Ellingson, E. 1999, ApJ, 527, 54 Barger, A. J., Arag´on-Salamanca, A., Ellis, R. S., Couch, W. J., Smail, I., & Sharples, R. 1996, MNRAS, 279, 1 Bekki, K. 1998, ApJ, 502, L133 Bekki, K., Shioya, Y., & Tanaka, I. 1999, ApJ, 520, L99 Bruzual A., G., & Charlot, S. 1993, ApJ, 405, 538 Butcher, H., & Oemler, A., Jr. 1978, ApJ, 219, 18 Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245 Couch, W. J., & Sharples, R. M. 1987, MNRAS, 229, 423 Couch, W. J., Ellis, R. S., Sharples, R. M., & Smail, I. 1994, ApJ, 430, 121 Couch, W. J., Barger, A. J., Smail, I., Ellis, R. S., & Sharples, R. M. 1998, ApJ, 497, 188 Dressler, A., & Gunn, J. E. 1983, ApJ, 270, 7 Dressler, A., & Gunn, J. E. 1992, ApJS, 78, 1 Dressler, A., Smail, I., Poggianti, B. M., Butcher, H., Couch, W. J., Ellis, R. S., & Oemler, A., Jr. 1999, ApJS, 122, 51 -- 13 -- Fioc, M., & Rocca-Volmerange, B. 1997, A&A, 326, 950 Fisher, D., Fabricant, D., Franx, M., & van Dokkum, P. 1998, ApJ, 498, 195 Lavery, R. J., & Henry, J. P. 1986, ApJ, 304, L5 Leitherer, C., & Heckmann, T. M. 1995, apjs, 96, 9 Matteucci, F., & Tornamb`e, A. 1987, A&A, 185, 51 & 1988, A&A, 196, 341 Mihos, J. C., & Hernquist, L. 1994, ApJ, 425, L13 Morris, S. L., Hutchings, J. B., Carlberg, R. G., Yee, H. K. C., Ellingson, E., Balogh, M. L., Abraham, R. G., & Smecker-Hane, T. A. 1998, ApJ, 507, 84 Owen, F. N., Ledlow, M. J., Keel, W. C., & Morrison, G. E. 1999, AJ, 118, 633 Planesas, P., Colina, L., & Perez-Olea, D. 1997, A&A, 325, 81 Poggianti, B. M., & Wu, H. 2000, ApJ, 529, 157 Poggianti, B. M., Smail, I., Dressler, A., Couch, W. J., Barger, J., Butcher, H., Ellis, E. S., & Oemler, A., Jr. 1999, ApJ, 518, 576 Shioya, Y., & Bekki, K. 1998, ApJ, 504, 42 Smail, I., Morrison, G., Gray, M. E., Owen, F. N., Ivison, R. J., Kneib, J.-P., & Ellis, R. S. 1999, ApJ, 525, 609 This manuscript was prepared with the AAS LATEX macros v4.0. -- 14 -- Fig. 1. -- The star formation history of our model. Starburst begins at Tsb = 7.64 Gyr (zsb = 0.4) and stops at Tend = 8.56 Gyr (zend = 0.29). We here assume that the mass of the model galaxy is 6 × 1010M⊙ (corresponding to the disk mass of the Galaxy). Fig. 2. -- Upper panel: The evolution of the equivalent width of [OII] emission line [EW([OII])]. The ND model (the model without dust extinction) and the SD one (with selective dust extinction) are shown as dotted line and solid line, respectively. Lower panel: The evolution of the equivalent width of Hδ [EW(Hδ)]. Red lines, blue lines, and black lines mean the equivalent width of the stellar absorption, the equivalent width of emission lines form ionized gas, and the sum of both lines, respectively. Positive (negative) EW(Hδ) in black lines means that Hδ line is observed as absorption (emission) line. Fig. 3. -- The evolution of galaxies on the EW([OII]) - EW(Hδ) plane. The ND model and the SD one are shown as a dotted line and solid one, respectively. Some points (T=1, 7.64, 8.56, 8.58, 12 Gyr) are shown as open circles (the ND model) and filled circles (the SD model) with age. The criteria of classification by Dressler et al. (1999) are also superimposed. Note that only the SD model can pass through the e(a) region during starburst (7.64 Gyr < T < 8.56 Gyr), which implies that the selective dust extinction during starburst is very important for the formation of e(a) galaxies. 40 20 ) r y / M ( R F S 0 0 5 10 Age (Gyr) ) A ( ) ] I I O [ ( W E ) A ( ) δ H ( W E 150 100 50 0 10 10 10 5 5 5 0 0 0 -- 5 -- 5 -- 5 -- 10 -- 10 -- 10 0 0 0 5 5 5 10 10 10 Age (Gyr) 100 100 e(b) 1 Gyr T = 7.64 Gyr sb 12 Gyr ) A ( ) ] I I O [ ( W E 10 10 e(c) 1 1 k -- 5 -- 5 end T = 8.56 Gyr e(a) 8.58 Gyr k+a a+k 0 0 5 5 10 10 EW (H ) (A) δ
astro-ph/0411802
1
0411
2004-11-30T17:20:46
Nearby early-type galaxies with ionized gas.I. Line-strength indices of the underlying stellar population
[ "astro-ph" ]
With the aim of building a data-set of spectral properties of well studied early-type galaxies showing emission lines, we present intermediate resolution spectra of 50 galaxies in the nearby Universe. The sample, which covers several of the E and S0 morphological sub-classes, is biased toward objects that might be expected to have ongoing and recent star formation, at least in small amounts, because of the presence of the emission lines. The emission are expected to come from the combination of active galactic nuclei and star formation regions within the galaxies. Sample galaxies are located in environments corresponding to a broad range of local galaxy densities, although predominantly in low density environments. Our long-slit spectra cover the 3700 - 7250 A wavelength range with a spectral resolution of about 7.6 A at 5550 A. The specific aim of this paper, and our first step on the investigation, is to map the underlying galaxy stellar population by measuring, along the slit, positioned along the galaxy major axis, line--strength indices at several, homogeneous galacto-centric distances. For each object we extracted 7 luminosity weighted apertures corrected for the galaxy ellipticity and 4 gradients and we measured 25 line-strength indices. The paper introduces the sample, presents the observations, describes the data reduction procedures, the extraction of apertures and gradients, the determination and correction of the line--strength indices, the procedure adopted to transform them into the Lick-IDS System and the procedures adopted for the emission correction. We finally discuss the comparisons between our dataset and line-strength indices available in the literature.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. AA3529rev2 (DOI: will be inserted by hand later) January 10, 2018 Nearby early -- type galaxies with ionized gas.I. Line-strength indices of the underlying stellar population⋆ Rampazzo R.1, Annibali F.2, Bressan A.1 , 2, Longhetti M.3, Padoan F.3, Zeilinger W.W.4, 1 INAF Osservatorio Astronomico di Padova, vicolo dell'Osservatorio 5, 35122 Padova, Italy 2 SISSA, via Beirut 4 - 34014 Trieste - Italy 3 INAF Osservatorio Astronomico di Brera, via Brera 28, I-20121 Milano, Italy 4Institut fur Astronomie der Universitat Wien, Turkenschanzstrasse 17, A-1180 Wien, Austria Received date; accepted date Abstract. With the aim of building a data-set of spectral properties of well studied early-type galaxies showing emission lines, we present intermediate resolution spectra of 50 galaxies in the nearby Universe. The sample, which covers several of the E and S0 morphological sub-classes, is biased toward objects that might be expected to have ongoing and recent star formation, at least in small amounts, because of the presence of the emission lines. The emission are expected to come from the combination of active galactic nuclei and star formation regions within the galaxies. Sample galaxies are located in environments corresponding to a broad range of local galaxy densities, although predominantly in low density environments. Our long -- slit spectra cover the 3700 - 7250 A wavelength range with a spectral resolution of ≈7.6 A at 5550 A. The specific aim of this paper, and our first step on the investigation, is to map the underlying galaxy stellar population by measuring, along the slit, positioned along the galaxy major axis, line -- strength indices at several, homogeneous galacto-centric distances. For each object we extracted 7 luminosity weighted apertures (with radii: 1.5′′, 2.5′′, 10′′, re/10, re/8, re/4 and re/2) corrected for the galaxy ellipticity and 4 gradients (0 ≤ r ≤re/16, re/16 ≤ r ≤re/8, re/8 ≤ r ≤re/4 and re/4 ≤ r ≤re/2). For each aperture and gradient we measured 25 line -- strength indices: 21 of the set defined by the Lick-IDS "standard" system (Trager et al. 1998) and 4 introduced by Worthey & Ottaviani (1997). Line -- strength indices have been transformed to the Lick-IDS system. Indices derived then include Hβ, Mg1, Mg2, Mgb, MgFe, Fe5270, Fe5335 commonly used in classic index-index diagrams. The paper introduces the sample, presents the observations, describes the data reduction procedures, the ex- traction of apertures and gradients, the determination and correction of the line -- strength indices, the procedure adopted to transform them into the Lick-IDS System and the procedures adopted for the emission correction. We finally discuss the comparisons between our dataset and line-strength indices available in the literature. A significant fraction, about 60%, of galaxies in the present sample has one previous measurement in the Lick -- IDS system but basically restricted within the re/8 region. Line-strength measures obtained both from apertures and gradients outside this area and within the re/8 region, with the present radial mapping, are completely new. Key words. Galaxies: elliptical and lenticular, cD -- Galaxies: fundamental parameters -- Galaxies: formation -- Galaxies: evolution 4 0 0 2 v o N 0 3 1 v 2 0 8 1 1 4 0 / h p - o r t s a : v i X r a 1. Introduction Ellipticals (Es) are among the most luminous and mas- sive galaxies in the Universe. Together with lenticular (S0) galaxies, composed of a bulge, a stellar disk and often a stellar bar component they form the vast category of early- type galaxies. Although, Es appear as a uniform class of galaxies, populating a planar distribution (the so-called Fundamental Plane) in the logarithmic parameter space defined by the central stellar velocity dispersion σ, the ef- Send offprint requests to: R. Rampazzo Correspondence to: [email protected] fective radius re and effective surface brightness Ie (see e.g. Djorgovski & Davis 1987), much evidence suggests that a secondary episode of star formation has occurred dur- ing their evolutionary history. Simulations indicate that galaxy collisions, accretion and merging episodes are im- portant factors in the evolution of galaxy shapes (see e.g. Barnes 1996; Schweizer 1996) and can interfere with their passive evolution. This understanding of early-type galaxy formation has been enhanced by the study of interstel- lar matter. This component and its relevance in secular galactic evolution was widely neglected in early studies of early-type galaxies since they were for a long time con- 2 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. sidered to be essentially devoid of interstellar gas. In the last two decades, however, multi-wavelength observations have changed this picture and have detected the presence of a multi-phase Inter Stellar Medium (ISM): a hot (107 K) and a warm (104 K) phase coexist and possibly in- terplay in several giant ellipticals. A cool (≈ 10 K) phase, detected in HI and CO, is also often revealed in early-type galaxies (Bettoni et al. 2001 and reference therein). Unlike spiral galaxies, the bulk of the gas in ellipticals is heated to the virial temperature, emitting in X-rays, and only comparatively small quantities are detected in the warm and cool phase of the interstellar medium (Bregman et al. 1992). The amount of X-ray emitting gas is related to the optical luminosity of the galaxy (White & Sarazin 1991), while no relation is found between the properties of the other components and the galaxy stellar luminosity. This latter fact may indicate that some amounts of the gaseous material may indeed be of external origin. A pioneering spectroscopic study (Phillips et al. 1986), that examined the properties of a set of 203 southern E and S0 galax- ies, began to shed light on the physical condition of the ionized gas in early-type galaxies. On the grounds of their analysis of the [NII 6583]/Hα ratio they pointed out that, when line emission is present, it is confined to the nucleus and the properties of giant ellipticals are "indistinguish- able" from a LINER nucleus (Heckman 1980). When the ionized region is imaged using narrow band filters cen- tered at Hα+[NII 6583], it appears extended with mor- phologies ranging from regular, disk-like structures to fila- mentary structures (see e.g. Ulrich-Demoulin 1984; Buson et al. 1993; Zeilinger et al. 1996; Macchetto et al. 1996) of several kpc in radius. The gaseous disks appear to be generally misaligned with respect to the stellar body of the galaxy suggesting an external origin for most of the gaseous matter. This picture is also supported by the ob- servations of decoupled kinematics of gas and stars in a significant fraction of early-type galaxies (Bertola et al. 1992a). Notwithstanding the large amount of studies, our current understanding of the origin and the nature of the "warm" ionized gas in elliptical galaxies is still rather un- certain. The fact that the emission regions are always as- sociated with dust absorption, even in the brightest X-ray systems, seems to exclude "cooling flows" as the origin of the ionized gas (see Goudfrooij 1998). The main ion- ization mechanism, which does not seem to be powered by star formation, however remains uncertain. Ionization mechanisms suggested range from photoionization by old hot stars -- post -- AGB and/or AGB-Manqu´e type objects (Binette et al. 1994) -- or mechanical energy flux from elec- tron conduction in hot, X-ray emitting gas (Voit 1991). Also ionization by a non-thermal central source is consid- ered. The present paper is the first of a series presenting a study of early-type galaxies in the nearby Universe show- ing emission lines in their optical spectra. Our aim is to improve the understanding of the nature of the ion- ized gas in early-type galaxies by studying its physical conditions, the possible ionization mechanisms, relations with the other gas components of the ISM and the con- nection to the stellar population of the host galaxy. The adopted strategy is to investigate galaxy spectra of inter- mediate spectral resolution at different galactocentric dis- tances and to attempt the modeling of their stellar pop- ulations to measure emission line properties. The study of stellar populations of early-type galaxies is of funda- mental importance to the understanding of their evolu- tion by the measurement of the evolution of the spec- tral energy distribution with time (see e.g. Buzzoni et al. 1992; Worthey 1992; Gonz´alez 1993; Buzzoni et al. 1994; Worthey et al.1994; Leonardi & Rose 1996; Wothey & Ottaviani 1997; Trager et al. 1998; Longhetti et al. 1998a; Vazdekis 1999; Longhetti et al. 1999; Longhetti et al. 2000; Trager et al. 2000; Kuntschner et al. 2000; Beuing et al. 2002; Kuntschner et al. 2002; Thomas et al. 2003; Mehlert et al. 2003) . Investigating issues such as the evo- lution of stellar populations and the ISM, we will explore the complex, evolving ecosystem within early -- type galax- ies and build a database of well studied galaxies to be used as a reference set for the study of intermediate and distant objects. Our target is to characterize the stellar popula- tions, in particular those related to the extended emission region, in order to constrain hints about the galaxy for- mation/evolution history from the modeling of the com- plete (lines and continuum) spectrum characteristics. In this paper we present the sample, the observations and the data reduction and we discuss, through the compari- son with the literature, the database of line -- strength in- dices we have measured. Forthcoming papers will analyze the emission region by tracing the properties as a func- tion of the distance from the galaxy center. The paper is organized as follows. Section 2 introduces the sample and some of the relevant properties useful to infer the ionized gas origin and nature. Section 3 presents the observations, the data reduction and the criteria and the methods used for the selection of apertures and gradients extracted from the long slit spectra. Section 4 details the transformation of the line-strength indices to the Lick-IDS System and provides the database of line -- strength indices measured at different galactocentric distances. In Section 5 we re- view the results, providing the database of line -- strength indices measured at different galactocentric distances and discuss the comparison with the literature. The Appendix A provides a description/comments of individual galaxies in the sample. 2. Characterization of the sample Our sample contains 50 early -- type galaxies. The sam- ple is selected from a compilation of galaxies showing ISM traces in at least one of the following bands: IRAS 100 µm, X-ray, radio, HI and CO (Roberts et al. 1991). All galaxies belong to Revised Shapley Ames Catalog of Bright Galaxies (RSA) (Sandage & Tammann 1987) and have a redshift of less than 5500 km s−1. The sam- ple should then be biased towards objects that might be expected to have ongoing and recent star formation, Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 3 Fig. 1. Distribution of B-magnitudes (first panel), morphological types (second panel), heliocentric velocity (third panel) and galaxy density (forth panel). 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 4000 5000 6000 7000 Fig. 2. Spectra of a representative galaxy in the sample. The figure shows the gradients obtained sampling each long -- slit spectrum in four regions: between 0 ≤ r ≤re/16 (indicated as "nuclear" in the figure), re/16 ≤ r ≤re/8, re/8 ≤ r ≤re/4 and re/4 ≤ r ≤re/2. For each region the figure overplots the two opposite sides of the galaxies with respect to the nucleus. With the exclusion of few cases (see text), the two sides agree within few percent. The major difference between the two sides often reside in the emission line distribution which is not symmetric with respect to the nucleus. 4 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. ident NGC 128 NGC 777 NGC 1052 NGC 1209 NGC 1297 NGC 1366 NGC 1380 NGC 1389 NGC 1407 NGC 1426 NGC 1453 NGC 1521 NGC 1533 NGC 1553 NGC 1947 NGC 2749 NGC 2911 NGC 2962 NGC 2974 NGC 3136 NGC 3258 NGC 3268 NGC 3489 NGC 3557 NGC 3607 NGC 3962 NGC 4552 NGC 4636 NGC 5077 NGC 5328 NGC 5363 NGC 5846 NGC 5898 NGC 6721 NGC 6868 NGC 6875 NGC 6876 NGC 6958 NGC 7007 NGC 7079 NGC 7097 NGC 7135 NGC 7192 NGC 7332 NGC 7377 IC 1459 IC 2006 IC 3370 IC 4296 IC 5063 Table 1 Overview of the observed sample R.A. (2000) Dec. 00 29 15.1 02 00 14.9 02 41 04.8 03 06 03.0 03 19 14.2 03 33 53.7 03 36 27.3 03 37 11.7 03 40 11.8 03 42 49.1 03 46 27.2 04 08 18.9 04 09 51.9 04 16 10.3 05 26 47.5 09 05 21.4 09 33 46.1 09 340 53.9 09 42 33.2 10 05 47.9 10 28 54.1 10 30 00.6 11 00 18.3 11 09 57.5 11 16 54.3 11 54 40.1 12 35 39.8 12 42 50.0 13 19 31.6 13 52 53.6 13 56 07.1 15 06 29.2 15 18 13.6 19 00 50.4 20 09 54.1 20 13 12.4 20 18 19.0 20 48 42.4 21 05 28.0 21 32 35.1 21 40 13.0 21 49 45.5 22 06 50.3 22 37 24.5 22 47 47.4 22 57 10.6 03 54 28.5 12 27 38.0 13 36 39.4 20 52 02.4 02 51 50 31 25 46 -08 15 21 -15 36 40 -19 05 59 -31 11 39 -34 58 34 -35 44 45 -18 34 48 -22 06 29 -03 58 08 -21 03 07 -56 07 07 -55 46 51 -63 45 38 18 18 49 10 09 08 05 09 57 -03 41 55 -67 22 41 -35 36 22 -35 19 32 13 54 05 -37 32 22 18 03 10 -13 58 31 12 33 23 02 41 17 -12 39 26 -28 29 16 05 15 20 01 36 21 -24 05 51 -57 45 28 -48 22 47 -46 09 38 -70 51 30 -37 59 50 -52 33 04 -44 04 00 -42 32 14 -34 52 33 -64 18 56 23 47 54 -22 18 38 -36 27 44 -35 57 58 -39 20 17 -33 58 00 -57 04 09 RSA S02(8) pec E1 E3/S0 E6 S02/3(0) E7/S01(7) S03(7)/Sa S01(5)/SB01 E0/S01(0) E4 E0 E3 SB02(2)/SBa S01/2(5)pec S03(0) pec E3 S0p or S03(2) RSB02/Sa E4 E4 E1 E2 S03/Sa E3 S03(3) E1 S01(0) E0/S01(6) S01/2(4) E4 [S03(5)] S01(0) S02/3(0) E1 E3/S02/3(3) S0/a(merger) E3 R?S01(3) S02/3/a SBa E4 S01 pec S02(0) S02/3(8) S02/3/Sa pec E4 E1 E2 pec E0 S03(3)pec/Sa RC3 S0 pec sp E1 E4 E6: SAB0 pec: S0 sp SA0 SAB(s)0-: E0 E4 E2 E3 SB0- SA(r)0 S0- pec E3 SA(s)0: pec RSAB(rs)0+ E4 E: E1 E2 SAB(rs)+ E3 SA(s)0: E1 E E0-1 E3+ E1: I0: E0+ E0 E+: E2 SAB(s)0- pec: E3 E+ SA0-: SB(s)0 E5 SA0- pec E+: S0 pec sp SA(s)0+ E E E2+ E SA(s)0+: P.A. 1 155 120 80 3 2 7 30 35 111 45 10 151 150 119 60 140 3 42 40 76 71 70 30 120 15 92 150 11 87 135 1 30 155 86 22 80 107 2 82 20 47 - 155 101 40 - 45 40 116 B0 12.63 12.23 11.53 12.26 12.61 12.81 11.10 12.39 10.93 12.37 12.59 12.58 11.65 10.36 11.75 13.03 12.53 12.71 11.68 11.42 12.48 12.57 11.13 11.23 11.08 11.61 10.80 10.50 12.52 12.78 11.06 11.13 12.41 12.93 11.72 12.66 12.45 12.13 12.92 12.49 12.48 12.61 12.15 11.58 12.61 10.96 12.27 11.91 11.43 13.14 (B-V)0 0.87 0.99 0.89 0.90 0.90 0.89 0.89 0.80 0.92 0.86 0.89 0.85 0.93 0.87 0.91 1.00 0.89 0.60 0.94 0.94 0.74 0.86 0.88 0.89 0.95 0.89 0.98 0.73 0.90 0.93 0.95 0.87 0.91 0.82 0.91 0.82 0.91 0.77 0.88 0.91 0.88 0.76 0.91 0.90 0.90 0.91 0.90 0.91 (U-B)0 0.51 0.44 0.40 0.44 0.39 0.34 0.56 0.42 0.46 0.41 0.44 0.44 0.43 0.51 0.23 0.37 0.41 0.34 0.48 0.43 0.54 0.46 0.54 0.50 0.55 0.45 0.52 0.29 0.57 0.40 0.40 0.25 0.42 0.45 0.48 0.25 0.29 0.54 0.42 0.35 0.54 0.26 Vhel 4227 5040 1475 2619 1550 1310 1844 986 1766 1443 3906 4165 773 1280 1100 4180 3131 2117 1890 1731 2778 2818 693 3038 934 1822 322 937 2764 4671 1138 1709 2267 4416 2854 3121 3836 2652 2954 2670 2404 2718 2904 1207 3291 1659 1350 2934 3762 3402 re 17.3 34.4 33.7 18.5 28.4 10.6 20.3 15.0 70.3 25.0 25.0 25.5 30.0 65.6 32.1 33.7 50.9 23.3 24.4 36.9 30.0 36.1 20.3 30.0 43.4 35.2 29.3 88.5 22.8 22.2 36.1 62.7 22.2 21.7 33.7 11.7 43.0 19.8 15.4 19.8 18.1 31.4 28.6 14.7 36.9 34.4 28.6 38.6 41.4 26.7 ρxyz 0.49 0.13 0.71 0.16 1.54 1.50 0.42 0.66 0.89 0.97 0.24 0.15 0.26 0.11 0.72 0.69 0.39 0.28 0.34 0.32 2.97 1.33 0.23 0.28 0.84 0.23 0.47 0.12 0.14 0.19 0.26 0.32 0.28 0.12 0.28 0.12 0.20 ǫ 0.67 0.21 0.28 0.52 0.13 0.56 0.41 0.37 0.07 0.34 0.17 0.35 0.19 0.38 0.11 0.07 0.32 0.37 0.38 0.24 0.13 0.24 0.37 0.21 0.11 0.22 0.09 0.24 0.15 0.31 0.34 0.07 0.07 0.15 0.19 0.41 0.13 0.15 0.42 0.32 0.29 0.31 0.15 0.75 0.19 0.28 0.15 0.21 0.17 0.28 Notes: The value of re of NGC 1297, NGC 6876 have been derived from ESO-LV (Lauberts & Valentijn (1989). at least in small amounts, because of the presence of emission lines. The emission should come from a com- bination of active galactic nuclei and star formation re- gions within the galaxies. Table 1 summarizes the ba- sic characteristics of the galaxies available from the lit- erature. Column (1) provides the identification; column (2) and (3) the R.A. & Dec. coordinates; column (4) and (5) the galaxy morphological classifications accord- ing to the RSA (Sandage & Tamman 1987) and RC3 (de Vaucouleurs et al. 1991) respectively. Columns (6), (7), (8) (9) give the position angle of the isophotes along major axis, the total corrected magnitude and the total (B-V) and (U-B) corrected colors from RC3 respectively. The heliocentric systemic velocity from HYPERCAT (http://www-obs.univ-lyon1.fr/hypercat) is reported in column (10). The effective radius, derived from Ae, the diameter of the effective aperture from RC3, is given in column (11). A measure of the richness of the environ- ment, ρxyz (galaxies Mpc−3), surrounding each galaxy is reported in column (12) (Tully 1988). Column (13) lists the average ellipticity of the galaxy as obtained from HYPERCAT. Figure 1 summarizes the basic characteristics of the present sample, and, in particular in the fourth panel, provides evidence that a large fraction of galaxies are in low density environments. In the following subsection we summarize morphological and photometric studies of the ionized component which provide an insight of the overall galaxy structure. In the Appendix A we complement the above information with individual notes on galaxies em- phasizing kinematical studies of the ionized gas compo- nent, its correlation with the stellar body and its possible origin. 2.1. Imaging surveys of the ionized gas component Buson et al. (1993) presented Hα+[NII] imaging of a set of 15 nearby elliptical and S0 galaxies with extended optical emission regions. Nine are included in our sample, namely NGC 1453, NGC 1947, NGC 2974, NGC 3962, NGC 4636, NGC 5846, NGC 6868, NGC 7097 and IC 1459. In most of Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 5 Table 2. Observing parameters Date of Observations Observer Spectrograph Detector Pixel size (µm) Scale along the slit (′′/px−1) Slit length (′) Slit width (′′) Dispersion(A mm−1) Spectral Resolution (FWHM at 5200 A ) (A) Spectral Range (A) Seeing Range(FWHM) (′′) Standard stars Run 1 March 98 Zeilinger W. B & C grating #25 Loral 2K UV flooded 15 0.82 4.5 2 187 7.6 3550-9100 1.2-2 Feige 56 Run 2 September 98 Zeilinger W. B & C grating #25 Loral 2K UV flooded 15 0.82 4.5 2 187 7.6 3550-9100 1.0-2.0 ltt 1788, ltt 377 Table 3. Journal of galaxy observations ident. Run Slit PA [deg] texp [sec] ident. Run Slit PA [deg] NGC 128 NGC 777 NGC 1052 NGC 1209 NGC 1297 NGC 1366 NGC 1380 NGC 1389 NGC 1407 NGC 1426 NGC 1453 NGC 1521 NGC 1533 NGC 1553 NGC 1947 NGC 2749 NGC 2911 NGC 2962 NGC 2974 NGC 3136 NGC 3258 NGC 3268 NGC 3489 NGC 3557 NGC 3607 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 1 1 1 1 1 1 1 1 1 155 120 80 3 2 7 30 35 111 45 10 151 150 29 60 140 3 42 40 76 71 70 30 120 2×1800 NGC 3962 2×1800 NGC 4552 2×1800 NGC 4636 1×2400 NGC 5077 2×1800 NGC 5328 2×1800 NGC 5363 2×1800 NGC 5846 2×1800 NGC 5898 2×1800 NGC 6721 2×1800 NGC 6868 2×1800 NGC 6875 2×1800 NGC 6876 1×2400 NGC 6958 2×1800 NGC 7007 2×1800 NGC 7079 2×1800 NGC 7097 2×1800 NGC 7135 1×1800 NGC 7192 2×1800 NGC 7332 2×1800 NGC 7377 2×1800 2×1800 2×1800 2×1800 2×1800 IC 1459 IC 2006 IC 3370 IC 4296 IC 5063 1 1 1 1 1 1 1 1 1,2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 2 15 92 150 11 87 135 1 30 155 86 50 75 107 2 82 20 47 90 155 101 40 45 45 40 116 texp [sec] 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 4×1800 2×1800 3×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 2×1800 these galaxies the extended emission forms an inclined disk with ordered motions (Zeilinger et al. 1996). Furthermore the major axes of the stellar and the gaseous components appear frequently misaligned. In NGC 1453 the emission appears strongly decoupled from the stellar component and roughly aligned with the minor axis of the galaxy. In NGC 1947 the emission appears associated with a complex systems of dust lanes and shows several distinct knots. In NGC 2974 the ionized gas appears to lie in a funda- mentally regular elongated structure with some periph- eral fainter filaments (dust is also present). Buson et al. (1993) reported that for NGC 3962 the emission region consists of two distinct subsystems: an elongated cen- tral component strongly misaligned with both the major and minor axes of the stellar figure, and a peculiar ex- tended arm-like structure departing from the major axis of the internal disk, crossing the stellar body at an angle of 180◦. Some dust is also noted. NGC 4636 has a ring- like emitting region extending asymmetrically around the galaxy nucleus. NGC 5846 shows complex, filamentary emission-line morphology, including an arm-like feature and dusty patches. NGC 6868 shows an elongated emitting 6 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Table 4. Velocity dispersion values adopted in the correction of line -- strength indices σre/8 [km s−1] σre/4 [km s−1] σre/2 [km s−1] Ref. Ident. σre/8 [km s−1] σre/4 [km s−1] σre/2 [km s−1] Ref. Ident. NGC 128 NGC 777 NGC 1052 NGC 1209 NGC 1297 NGC 1366 NGC 1380 NGC 1389 NGC 1407 NGC 1426 NGC 1453 NGC 1521 NGC 1533 NGC 1553 NGC 1947 NGC 2749 NGC 2911 NGC 2962 NGC 2974 NGC 3136 NGC 3258 NGC 3268 NGC 3489 NGC 3557 NGC 3607 183 317 215 240 115 120 240 139 286 162* 289 235 174 180 142 248 235 197 220 230 271 227 129 265 220 -- 272 179 195 -- -- 220 -- -- 157 -- 236 -- 142 142 221 -- 168 170 180 313 155 116 247 210 -- 266 179 178 -- -- JS89 FI94 PS98 198 DO95 NGC 3962 NGC 4552 NGC 4636 NGC 5077 NGC 5328 NGC 5363 NGC 5846 NGC 5898 NGC 6721 NGC 6868 PS97a PS98 NGC 6875 NGC 6876 NGC 6958 L98,R88 NGC 7007 NGC 7079 BGZ92 NGC 7097 JS89 NGC 7135 NGC 7192 NGC 7332 NGC 7377 PS00 CvM94 KZ00 KZ00 KZ00 CMP00 FI94 CMP00 IC 1459 IC 2006 IC 3370 IC 4296 IC 5063 225 264 209 260 303 199 250 220 262 277 -- 230 223 125 155 224 231* 257 136 145 311 122 202 340 160 -- 226 202 239 -- 181 228 183 245 235 -- -- 171 -- 125 234 239 247 127 -- 269 -- 146 310 -- -- 214 228 -- 148 190 172 171 220 -- -- 137 -- 85 196 -- 266 116 -- 269 -- 127 320 -- SP97b CMP00 CMP00 S83 CMP00 CMP00 B94 CMP00 L98 BG97 C86 L98 CD94 SP97c FI94 CDB93 J87 S93 -- -- 121 -- 206 -- -- 142 -- -- 119 130 264 -- 115 220 195 the central value, obtained average Notes: HYPERCAT (http://www-obs.univ-lyon1.fr/hypercat/), is adopted for σre/8. Values at σre/4 and σre/2 velocity dispersion values are obtained from references quoted in columns 5 and 10. Legend: DO95 = D'Onofrio et al. 1995; SP97a = Simien & Prugniel (1997a); SP97b = Simien & Prugniel (1997b); PS97 = Prugniel & Simien (1997); PS00 = Prugniel & Simien (2000); PS98 = Prugniel & Simien (1998); FI94 = Fried & Illingworth (1994); Z96 = Zeilinger et al. (1996); JS89 = Jedrzejewski & Schecther (1989); R88 = Rampazzo (1988); L98 = Longhetti et al. (1998b); BGZ92 = Bertola et al. (1992a); CMP00 = Caon et al. (2000); KZ00 = Koprolin & Zeilinger (2000); CDB93 = Carollo et al. (1993); S83 = Sharples et al. (1983); B94 = Bertin et al. (1994 ); V87 = Varnas et al. (1987); C86 = Caldwell et al. (1986); CD94 = Carollo & Danziger (1994); J87 = Jarvis (1987); BG97 = Bettoni & Galletta (1997); S93 = Saglia et al. (1993); CvM94 = Cinzano & van der Marel (1994). compilation from the on-line region, with faint peripheral extensions and dust patches, which are strongly decoupled from the stellar component. NGC 7079 has elongated emission misaligned by about 30◦ with the stellar figure associated with dusty features. The ionized region of IC 1459 is aligned with the stellar major axis. Dust features for this latter galaxy have been identified by Goudfrooij (1994). A characterization of the extended emission region was attempted by Macchetto et al. (1996) who observed 73 luminous early -- type galaxies selected from the RC3 catalogue. CCD images were ob- tained in broad R-band and narrow band images centered on the Hα + [NII] emission lines. A set of 17 galaxies of our sample galaxies are in the Macchetto et al. set, namely NGC 1407, NGC 1453, NGC 3489, NGC 3607, NGC 3268, NGC 4552, NGC 4636, NGC 5077, NGC 5846, NGC 5898, NGC 6721, NGC 6868, NGC 6875, NGC 6876, NGC 7192, IC 1459 and IC 4296. The morphology of the extended emission has been divided into three categories. "SD" in- dicates galaxies with a small disk, extended on average less than 4 kpc (they adopt H0=55 km s−1 Mpc−1) with sometimes faint and short filaments. The ionized regions of NGC 1407, NGC 3268, NGC 3489, NGC 3607, NGC 4552 and NGC 4636 were classified as "SD". "RE" indicates regular extended regions, similar to previous ones but ex- tended between 4 and 18 kpc. NGC 1453, NGC 5077, NGC 5846, NGC 5898, NGC 6721, NGC 6868, NGC 6875, NGC 7192 and IC 1459 have extended ionized regions. "F" represents the detection of filaments and conspicuous fil- amentary structure which dominate the morphology and which depart from a more regular disk -- like inner region. The filamentary structures extend as far as 10 kpc from the galaxy center. However, our sample contains no galax- ies with a filamentary morphology. Macchetto and co- authors did not classify the ionized regions of NGC 6876 and IC 4296. From the above notes it appears that the ionized region of a large fraction of galaxies in our sam- ple have a disk -- like structure. A large number of kine- matical studies (see individual notes in Appendix A) sup- Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 7 ports this hypothesis. In addition, it emerges that the ion- ized emission is always associated with dust absorption, even in the brightest X-ray systems (Goudfrooij 1994; Goudfrooij 1998). Several classes of galaxies are present in the sample: interacting or post-interacting galaxies, galaxies showing evidence of kinematical decoupling be- tween galaxy sub-components, elliptical galaxies with a dust lane along the minor axis, radio galaxies and galax- ies hosting an AGN nucleus. To summarize the individ- ual notes in Appendix A: (a) the sample contains four galaxies showing a shell structure (namely NGC 1553, NGC 4552, NGC 6958, NGC 7135). Twenty galaxies (namely NGC 128, NGC 1052, NGC 1407, NGC 1947, NGC 2749, NGC 3136, NGC 3489, NGC 4636, NGC 5077, NGC 5363, NGC 5846, NGC 5898, NGC 6868, NGC 7007, NGC 7097, NGC 7192, NGC 7332, IC 1459, IC 2006, IC 4296) have a peculiar kinematical behavior, i.e. rota- tion along the apparent minor axis, turbulent gas motions, counterrotation of star vs. gas and/or star vs. stars. In four galaxies (namely NGC 1052, NGC 1553, NGC 3962 and NGC 7332) multiple gas components have been detected. 3. Observations and data-reduction 3.1. Observations Galaxies were observed during two separate runs (March and September 1998) at the 1.5m ESO telescope (La Silla), equipped with a Boller & Chivens spectrograph and a UV coated CCD Fa2048L (2048×2048) camera (ESO CCD #39). Details of the observations and typical seeing con- ditions during each run are reported in Table 2. Table 3. provides a journal of observations i.e. the object identifi- cation (column 1,5), the observing run (columns 2,6), the slit position angle oriented North through East (columns 3,7) and the total exposure time (columns 4,8). The spec- troscopic slit was oriented along the galaxy major axis for most observations. He-Ar calibration frames were taken before and after each exposure to allow an accurate wave- length calibration to be obtained. 3.2. Data reduction Pre-reduction, wavelength calibration and sky subtraction were performed using the IRAF1 package. A marginal misalignment between CCD pixels and the slit has been checked and corrected on each image by means of an "ad hoc" written routine. The wavelength range covered by the observations was ≈ 3550 -- 9100 A. Fringing seriously affected observations longward of ≈ 7300 A. After accurate flat-fielding correction we con- sidered the wavelength range 3700 - 7250 Afor further use. Multiple spectra for a given galaxy were co-added. Relative flux calibration was obtained using a sequence 1 IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy Inc., under coopera- tive agreement with the National Science Foundation of spectrophotometric standard stars. Before flux calibra- tion, frames were corrected for atmospheric extinction, tai- lored to the ESO La Silla coefficients. The redshift value of each galaxy was directly measured from the lines of spec- tra. Spectra were finally de-redshifted to the rest frame. A set of representative spectra of the galaxies in the sam- ple are presented in Figure 2. The figure also shows the similarity of the two sides of the galaxy with respect to the nucleus: surprisingly after the geometrical and red- shift corrections the two sides compare within few (2-3) percent, the major deviations due to the asymmetric dis- tribution of the emission within the galaxies. For some galaxies, namely NGC 1947, NGC 2911, NGC 5328 and NGC 6875 there are serious differences between the two sides with respect to the nucleus since the spectrum is contaminated by the presence of a foreground star (see also next section). 3.3. Extraction of apertures and gradients We have extracted flux-calibrated spectra along the slit in seven circular concentric regions, hereafter "aper- tures", and in four adjacent regions, hereafter "gradients". Aperture spectra were sampled with radii of 1.5′′, 2.5′′, 10′′, re/10, re/8, re/4 and re/2. Our aperture spectra are suitable for comparison with typical apertures commonly used in the literature both with different galaxies sam- ple and with ongoing galaxy surveys (e.g. the SLOAN fiber spectra). The apertures were simulated by assuming that each radial point along the semi-major axis (sampled at both sides of our slit) is representative of the corre- sponding semi-ellipsis in the two-dimensional image. The galaxy ellipticity, ε is given in Table 1 col. (13) and has been assumed to be constant with radius. With this triv- ial relation between a point within the simulated semi- circular apertures and the spectrum along the semi-major axis on the same side, we have calculated the average sur- face brightness spectrum and the corresponding luminos- ity weighted radius, rl, of each semi-circular aperture. The average radius and the flux in each aperture are given by the formulae: hrli = R rFλ(r, ε)ds R Fλ(r, ε)ds Fλ = R Fλ(r, ε)ds R ds (1) (2) where r, ε and s are the radius along the slit, the ellip- ticity and the area respectively. This procedure allow us to obtain a fair estimate of the aperture measurement cor- responding to a mono-dimensional spectrum, in the lack of spectra along the minor axis (see e.g. Gonz´alez, 1993 for a thorough discussion). Simulated aperture spectra sample an increasing con- centric circular region. However, the S/N of our spectra is enough to obtain information on the spatial gradients. To this purpose we have extracted spectra also in four ad- jacent regions along each semi-major axis, 0 ≤ r ≤re/16 8 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. ("nuclear"), re/16 ≤ r ≤re/8, re/8 ≤ r ≤re/4 and re/4 ≤ r ≤re/2) providing the linear average flux in the above interval. The average radius and the flux in each interval are given by the formulae: hrli = R rFλ(r)dr R Fλ(r)dr Fλ = R Fλdr R dr (3) (4) Figure 2 shows the spectra of the gradient of IC 1459 as a representative galaxy. Each panel displays the spec- tra extracted from the two symmetric sides with respect to the nucleus. The strategy of averaging the two sides of the spectrum with respect to the nucleus deserves few comments. Each single galaxy demonstrates to be quite homogeneous because in general the variations of the op- posite sides are well within a few (2-3) percent in most of the cases. Even very faint features are well replicated in each side suggesting both that they are real photospheric features and a radial homogeneity of the stellar popula- tion. In few cases, we notice that the emission features are less prominent (or even absent) in one side of the galaxy with respect to the other. The ionized gas component is less homogeneously distributed than the stellar compo- nent. This could suggest that the gas is not in an equilib- rium configuration in the potential well of the galaxy and possibly accreted from outside. An alternative explanation could be that the excitation mechanism is local (see also Appendix A). The study of the emission features as func- tion of the distance from the galaxy center will be deal with in a forthcoming paper. Given the homogeneity of the side-spectra we present indices for the averaged spectra. However, due to the con- tamination of the spectrum by foreground stars we mea- sured for NGC 1947, NGC 2911 and NGC 6875 apertures and gradients up to r≤re/8, while for NGC 5328 we con- sider apertures and gradients up to r≤re/4. 4. Measurements of line-strength indices and transformation to the Lick-IDS System In the following sub-sections we detail the procedure we adopted to extract line -- strength indices from the original spectra and to transform them into the Lick -- IDS System. We measured 21 line-strength indices of the original Lick- IDS system using the redefined passbands (see Table 2 in Trager et al. 1998 for the index definitions) plus 4 new line strength indices introduced by Worthey & Ottaviani (1997) (see their Table 1 for the index definitions and Table 2 in Trager et al. 1998). In the subsequent analysis we then derived this set of 25 indices. We tested our index- measuring pipeline on the original Lick spectra comparing our measurements with the index values given by Worthey http://astro.wsu.edu/worthey/html/system.html. Fig. 3. Comparison between central velocity dispersions used in this paper and in Trager et al. (1998). The lines indicate an average error (20 km s−1) in the central veloc- ity dispersion measurements. Table 5. Lick standard stars Ident. Spectral Type Numb. obser.s HD 165195 HD 172401 HD 23430 HR 6159 HR 6710 HR 6770 HR 6868 HR 7317 HR 3145 HR 3418 HR 3845 HR 4287 HR 5480 HR 5582 HR 5690 HR 5888 HR 6299 K3p K0III A0 K4III F2IV G8III M0III K4III K2III K1III K2.5III K1III G7III K3III K5III G8III K2III 1 1 2 2 1 1 1 2 3 3 1 2 1 2 1 1 1 4.1. Spectral resolution Our spectral resolution (FWHM ∼ 7.6 A at ∼ 5000 A) on the entire spectrum is slightly larger than the wavelength-dependent resolution of the Lick -- IDS system (see Worthey & Ottaviani 1997). In order to conform our measures to the Lick-IDS System, we smoothed our data convolving each spectrum (apertures and gradients) with Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 9 a wavelength-dependent gaussian kernel with widths given by the formula: Table 6. Hβ corrections for apertures σsmooth(λ) = r F W HM (λ)2 8 ln 2 Lick − F W HM (λ)2 our (5) The selection of a gaussian kernel is justified by the gaussian shape of both our and Lick spectra (Worthey & Ottaviani 1997) absorption lines. 4.2. Correction for velocity dispersion The observed spectrum of a galaxy can be regarded as a stellar spectrum (reflecting the global spectral charac- teristics of the galaxy) convolved with the radial velocity distribution of its stellar population. Therefore spectral features in a galactic spectrum are not the simple sum of its corresponding stellar spectra, because of the stellar motions. To measure the stellar composition of galaxies, we need to correct index measurements for the effects of the galaxy velocity dispersion (see e.g. G93, Trager et al. (1998); Longhetti et al. (1998a)). To this purpose, among the Lick stars observed to- gether with the galaxies (see also Section 4.4), we have selected stars with spectral type between G8III and K2III (7 stars in our sample) typically used as kinematical tem- plates in early-type galaxies. The list of the observed stars, as well as their spectral type, is given in Table 5. The stel- lar spectra (degraded to the Lick resolution) have been convolved with gaussian curves of various widths in or- der to simulate different galactic velocity dispersions. We have considered a grid of velocity dispersion values in the range (80 − 350)kms−1. The values of velocity disper- sion adopted for line-strength correction in the present paper are in agreement with those adopted by Trager et al. (1998), as shown in Figure 3, and well within the above velocity dispersion range . On each convolved spectrum we have measured the 25 Lick-indices. The fractional index variations have been derived for each velocity dispersion, σ, of our grid through an average on the selected stellar spectra: Ri,σ = 1 N N ( Xj=1 EWi,j,σ − EWi,j,0 EWi,j,0 ) (6) where N is the number of studied stars, EWi,j,σ is the i-th index measured on the j-th star at the velocity dispersion σ, and EWi,j,0 is the measured index at zero velocity dispersion. To compute the corrections for velocity dispersion, we derived at the radius of each aperture and gradient the corresponding σ value using the data listed in Table 4. The tabulated values characterize the trend of each galaxy velocity dispersion curve. For galaxies having only the cen- tral (re/8) value of σ we use this value also for the cor- rection of the indices at larger radii (the tables of indices uncorrected for velocity dispersion are available, under re- quest, to authors which may apply suitable corrections Galaxy NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 ........ aperture EWO[III] Quality EWHα -0.558 -0.532 -0.481 -0.465 -0.325 -0.301 -0.186 -0.126 -0.114 -0.100 -0.103 -0.113 -0.124 -0.124 ....... -0.263 -0.210 -0.205 -0.182 -0.053 0.006 0.036 0.039 0.077 0.060 0.076 0.112 0.144 0.071 ...... 2.000 1.000 1.000 1.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 ...... 0 1 2 3 4 5 6 0 1 2 3 4 5 6 ..... The table provides in column 2 the number of the aperture, the correspondent radius of which is given in Table 9, and the estimate of the EW of the O[III] and Hα (columns 3 and 5 respectively) obtained from the subtraction of the template galaxy, NGC 1426. Column 4 gives the quality of the measure, obtained as the ratio between the estimated emission (Fgal - Ftempl) and the variance of the spectrum in the O[III] wave- length region, σλ. When the emission is lower than the variance the quality is set to 0. When the emission is between 1 and 2 σλ or larger than 2 σλ the quality is set to 1 and 2 respectively. The entire table is given is electronic form. when new extended velocity dispersion curve measures will be available). The new index corrected for the effect of velocity dis- persion is computed in the following way: EWi,new = EWi,old/(1 + Ri,sigma) (7) where Ri,σ is determined by interpolation of the σ value on the grid of velocity dispersions. 4.3. Correction of the Hβ index for emission The presence of emission lines affects the measure of some line -- strength indices. In particular, the Hβ index mea- sure of the underlying stellar population could be con- taminated by a significant infilling due to presence of the Hβ emission component. Gonz´alez et al. (1993) verified a strong correlation be- tween the Hβ and the [OIII] emission in his sample, such that EW (Hβem)/EW ([OIII]λ5007) = 0.7. Trager et al. (2000) examined the accuracy of this correlation by study- ing the Hβ/[OIII] ratio supplementing the G93 sample with an additional sample of early-type galaxies with emis- sion lines from the catalog of Ho, Filipenko & Sargent (1997). They found that the Hβ/[OIII] ratio varies from 0.33 to 1.25, with a median value of 0.6. They propose that the correction to the Hβ index is ∆ Hβ = 0.6 EW([OIII]λ 5007). The first step in order to measure the EW([OIII]λ5007) of the emission line is to degrade each spectrum (apertures 10 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Fig. 4. (left four panels) Comparison between the correction estimate of the Hβ emission derived from [OIII] emission line and from the Hα emission (see Section 4.3). The solid line is the one-to-one relation assuming for O[III] valid the formula EW (Hβem)/EW ([OIII]λ5007) = 0.7. The comparison is shown in the four regions sampled by the linear gradients (legend: lg1 (0 ≤ r ≤re/16("nuclear")), lg2 (re/16 ≤ r ≤re/8), lg3 (re/8 ≤ r ≤re/4) and lg4 (re/4 ≤ r ≤re/2). Open circles indicate galaxies which O[III] emission is detected under 1σ level, triangles and full squares between 1 and 2σ levels and above 2σlevel respectively (see text). (right four panels) Fitting of N[II](λ 6548, 6584) and Hα lines for two representative galaxies: one with Hα in emission (NGC 2749) and the second with the Hα infilling (IC 2006). Lower panels show the residuals lines after the subtraction of the Hα line of the template galaxy NGC 1426 (dotted lines in the upper panels). Table 7. Hβ corrections for gradients Galaxy NGC128 NGC128 NGC128 NGC128 NGC777 NGC777 NGC777 NGC777 .... 0 1 2 3 0 1 2 3 ... aperture EWO[III] Quality EWHα -0.702 -0.608 -0.434 -0.174 -0.124 -0.106 -0.168 -0.129 -0.317 -0.204 -0.173 0.119 0.042 0.057 0.045 0.162 ..... 1.000 1.000 2.000 0.000 0.000 0.000 0.000 0.000 ..... ..... The table provides in column 2 the gradient, (0=0 ≤ r ≤re/16, 1 = re/16 ≤ r ≤re/8, 2 = re/8 ≤ r ≤re/4 and 3 = re/4 ≤ r ≤re/2), and the estimate of the EW of the O[III] and Hα (columns 3 and 5 respectively) obtained from the subtraction of the template galaxy, NGC 1426. Column 4 gives the quality of the measure, obtained as the ratio between the estimated emission (Fgal - Ftempl) and the variance of the spectrum in the O[III] wavelength region, σλ. When the emission is lower than the variance the quality is set to 0. When the emission is between 1 and 2 σλ or larger than 2 σλ the quality is set to 1 and 2 respectively. The entire table is given in electronic form. and gradients) to the Lick resolution. The second, more delicate step, is to "build" a suitable template for the un- derlying stellar component and then adapt it to the galaxy velocity dispersion. At this purpose different methods have been adopted both using stellar and galaxy templates. Gonz´alez (1993) used stellar templates. The adopted technique adopted consists in simultaneously fitting the kinematics and the spectrum of a galaxy with a library of stellar spectra. However, we are aware that absorption line spectra of early-type galaxies cannot be adequately fitted using Galactic stars or star clusters, the main reason being the high metallicity in giant ellipticals and the non-solar element ratios in ellipticals. To compute the emissions Goudfrooij (1998) suggested to use a suitable template galaxy spectrum of an elliptical. Following his suggestion we then considered galaxies in our sample that, according to our observed spectrum and the combined information coming from the literature, show neither evidence in their spectrum of neither emission features nor of dust, usually associated with the gas emission (see Goudfrooij 1998), in their image. To this purpose we adopted NGC 1426 spec- trum as a template for an old population: this choice is motivated both by the lack of emission line in its spectrum and by the absence of dust features in high resolution HST images obtained by Quillen et al. (2000) (see Appendix A). We then proceeded in following way: we smoothed the template spectrum to adapt it to the velocity disper- Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 11 sion of the galaxy region under exam and normalized it to the galaxy continuum on both sides of Hα line. All spec- tra (aperture and gradients) have been analyzed using the template in the corresponding region. NGC 1426 has a low value of velocity dispersion consistently with its low Mg2 index; this indicates that it is not a giant elliptical and may be representative of the metal poor tail in our sam- ple. Given the anti-correlation between Hβ index and Mg2 index strength one may wonder whether this galaxy is the suitable template for all the sample. In order to check the reliability of the use of this template we have compared the Hα absorption profile of NGC 1426 with that of NGC 1407, which belongs to upper tail of the Mg2 σ relation. Once adopted the above procedure of smoothing and nor- malization we notice that the residual difference in the Hα profile implies a negligible difference in the computed EW Hβ correction (≃ 0.03 A). We characterized the emission as the flux in excess with respect to the template within the bandpass (4996.85 - 5016.85) centered at 5007A, while the continuum is de- fined by a blue (4885.00 - 4935.00) and a red (5030.00 - 5070.00) bandpass (Gonz´alez 1993): EWem = Z λ2 λ1 FR − Ftemp FC dλ (8) where FR, Ftemp and FC are the galaxy, the tem- plate and the continuum fluxes respectively. According to this definition, detected emissions result as negative EWs. Considering the ([OIII]λ5007) emissions detected above 1 σ (the variance of the spectrum), we derived the EW of the Hβ emission from the equation ∆ Hβ = 0.7 EW([OIII]λ 5007). The derived corrections for Hβ could be easily com- pared with Gonz´alez (1993) for the three galaxies in com- mon (namely NGC 1453, NGC 4552 and NGC 5846). We obtained an 0.48 (vs. 0.89±0.06), 0.02 (vs. 0.25±0.05) and 0.09 (vs. 0.39±0.08) i.e. systematically smaller cor- rections as if our template had a residual gas infilling, but which is not confirmed by imaging observations as out- lined above. We tested also the use of a stellar template taken from our observed Lick stars, the main problem be- ing that our set of stars is very "limited" with respect to that of Lick stellar library used by Gonz´alez. The use of stellar templates (K giant stars) in our sample, maintain- ing the above O[III] bandpasses, implies a worse match of the spectral features in our galaxy sample than if we adopt NGC 1426 as a template. Finally this results in a system- atically higher Hβ corrections, at least for the galaxies in common with Gonz´alez. The large wavelength coverage of our spectra permits us to measure also the Hα emission and allows a fur- ther estimate of the Hβ emission according to the relation FHβ = 1/2.86FHα (see e.g. Osterbrock 1989). The measure of the Hα emission is not straightfor- ward in our spectra since the line is blended with the ([NII]λ6548, 6584) emission lines. To derive the Hα emis- sions we fitted each galaxy spectrum (apertures and gra- dients) with a model resulting from the sum of our tem- Table 8 α and β coeff. for index correction Index α β aver. disp. CN1 CN2 Ca4227 G4300 Fe4383 ca4455 Fe4531 Fe4668 Hβ Fe5015 Mg1 Mg2 Mgb Fe5270 Fe5335 Fe5406 Fe5709 Fe5782 NaD TiO1 TiO2 HδA HγA HδF HγF 1.059 1.035 1.317 1.105 0.963 0.451 1.289 0.976 1.064 1.031 1.014 0.998 1.014 1.058 0.990 1.005 1.321 1.167 1.003 0.997 1.003 1.136 0.990 1.059 1.011 0.023 0.030 0.408 0.179 1.169 1.844 -0.299 0.128 -0.196 0.804 0.015 0.020 0.417 0.270 0.356 0.282 -0.270 -0.075 0.027 0.004 -0.001 -0.622 0.518 -0.036 0.458 0.025 0.023 0.396 0.310 0.772 0.341 0.437 0.653 0.166 0.396 0.009 0.012 0.241 0.240 0.249 0.151 0.174 0.165 0.245 0.006 0.008 1.087 0.734 0.503 0.745 unit mag mag A A A A A A A A mag mag A A A A A A A mag mag A A A A plate galaxy spectrum and three gaussian curves of arbi- trary widths and amplitudes (see Figure 4). Once derived the Hβ emitted fluxes from equation (8) we computed the pseudo-continua in Hβ according to the bandpass defini- tion of Trager et al. (1998) and used them to transform flux measures into EWs. In Figure 4 (left panels) we plot the comparison be- tween the two different estimates computed in the four gradients. The points in the figure do not include the Seyfert 2 galaxy IC 5063 for which the ∆Hβ as derived from the [OIII] emission is significantly higher than the value resulting from Hα emission. The new Hβ index, corrected for the emission in- filling, is computed from the non raw one according to the formula EW (Hβcorr) = EW (Hβraw) − Hβem, where EW(Hβcorr) is the corrected value obtained applying the Hβem estimate derived from the [OIII] emission using as template the galaxy NGC 1426. This latter estimate is statistically similar to that obtained from the Hα correc- tion as shown in Figure 4, although the use of Hα estimate for emission correction will be widely discussed in a forth- coming paper. Table 6 and Table 7 report the values of the Hβ correction for the apertures and gradients derived from EWO[III] and Hα (complete tables are given in electronic form). 12 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Fig. 5. Comparison of passband measurements from our spectra and original Lick data for 17 Lick standard stars. The dotted line is the one to one relation while the solid line is the robust straight-line fit to the filled squares. 4.4. Lick-IDS Standard Stars After the indices have been homogenized to the Lick-IDS wavelength dependent resolution, corrected for emission and velocity dispersion, we still have to perform a last step to transform our line-strengths indices into the Lick sys- tem. To this purpose we followed the prescription given by Worthey & Ottaviani (1997) and observed, contemporary to the galaxies, a sample of 17 standard stars of different spectral type common to the Lick library. The observed stars, together with the corresponding number of obser- vations and the spectral type, are given in Table 5. Once the stellar spectra have been degraded to the wavelength- dependent resolution of the Lick system, we have mea- sured the line-strength indices with the same procedure adopted for the galaxy spectra. We compared our mea- sures with the Lick-IDS indices reported by Worthey et al. 1994 for the standard stars. The deviations of our mea- sures from the Lick system are parametrized through a robust straight-line fit (see e.g. Numerical Recipes 1992) which avoid an undesired sensitivity to outlying points in two dimension fitting to a straight line. The functional form is EWLick = β + α × EWour where EWour and EWLick are respectively our index measure on the stel- lar spectrum and the Lick value given in Worthey et al. 1994. Fig. 5 shows the comparison between the Lick indices and our measures for the observed standard stars. The dotted line represent the one to one relation while the solid line is the derived fit. For each index we report the coefficients α and β of the fit in Table 8. Notice that for Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. galaxy galaxy NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 ... Table 9 Fully corrected line -- strength indices for the apertures iz iz 0 0 1 1 2 2 3 3 4 4 5 5 6 6 0 0 1 1 2 2 3 3 4 4 5 5 6 6 ... ie ie 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 ... r1 rl 0.000 0.055 0.000 0.062 0.000 0.076 0.000 0.087 0.000 0.142 0.000 0.161 0.000 0.260 0.000 0.028 0.000 0.045 0.000 0.059 0.000 0.072 0.000 0.081 0.000 0.125 0.000 0.210 ... r2 re 0.087 17.300 0.100 17.300 0.125 17.300 0.145 17.300 0.250 17.300 0.289 17.300 0.500 17.300 0.044 34.400 0.073 34.400 0.100 34.400 0.125 34.400 0.145 34.400 0.250 34.400 0.500 34.400 ... CN1 eCN1 0.136 0.003 0.136 0.003 0.135 0.003 0.135 0.003 0.132 0.003 0.131 0.003 0.125 0.003 0.176 0.002 0.175 0.004 0.173 0.004 0.167 0.002 0.165 0.002 0.159 0.002 0.174 0.005 ... CN2 eCN2 0.167 0.004 0.166 0.004 0.165 0.004 0.164 0.004 0.159 0.004 0.158 0.004 0.151 0.004 0.215 0.003 0.212 0.004 0.209 0.004 0.205 0.003 0.203 0.003 0.194 0.003 0.199 0.005 ... Ca4227 eCa4227 1.830 0.054 1.830 0.054 1.830 0.054 1.840 0.055 1.760 0.056 1.770 0.056 1.760 0.059 2.410 0.046 2.050 0.055 1.870 0.053 1.870 0.046 1.870 0.045 1.860 0.046 1.920 0.074 ... G4300 eG4300 6.150 0.087 6.150 0.087 6.150 0.087 6.150 0.087 6.120 0.090 6.120 0.091 6.090 0.095 6.030 0.083 6.100 0.104 6.120 0.101 6.160 0.080 6.160 0.080 6.250 0.077 6.440 0.130 ... Fe4383 eFe4383 5.790 0.149 5.850 0.149 5.950 0.150 5.980 0.151 5.910 0.154 5.920 0.155 5.970 0.160 5.050 0.131 4.940 0.174 4.980 0.168 5.140 0.127 5.160 0.127 5.070 0.123 5.640 0.216 ... Ca4455 eCa4455 2.330 0.079 2.340 0.080 2.350 0.080 2.350 0.080 2.300 0.082 2.300 0.083 2.300 0.086 2.290 0.070 2.290 0.090 2.340 0.086 2.400 0.065 2.440 0.065 2.470 0.064 2.530 0.108 ... Fe4531 eFe4531 4.150 0.139 4.150 0.139 4.160 0.141 4.150 0.141 4.090 0.144 4.100 0.145 4.130 0.152 3.910 0.122 4.030 0.163 4.050 0.158 3.990 0.115 3.940 0.115 3.820 0.115 3.960 0.204 ... The complete Table 9 is given in electronic form. Table 10 Fully corrected line -- strength indices for the gradients galaxy galaxy NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC128 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 NGC777 ... iz iz 0 0 1 1 2 2 3 3 0 0 1 1 2 2 3 3 ... ie ie 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 ... r1 rl 0.000 0.031 0.063 0.093 0.125 0.182 0.250 0.359 0.000 0.029 0.063 0.090 0.125 0.177 0.250 0.347 ... r2 re 0.063 17.300 0.125 17.300 0.250 17.300 0.500 17.300 0.063 34.400 0.125 34.400 0.250 34.400 0.500 34.400 ... CN1 eCN1 0.140 0.004 0.137 0.003 0.130 0.003 0.138 0.003 0.175 0.003 0.170 0.003 0.146 0.003 0.189 0.004 ... CN2 eCN2 0.167 0.005 0.168 0.004 0.160 0.004 0.162 0.004 0.213 0.004 0.206 0.003 0.180 0.004 0.210 0.005 ... Ca4227 eCa4227 1.580 0.077 1.790 0.053 1.800 0.054 1.850 0.057 2.280 0.063 1.670 0.047 1.760 0.056 2.040 0.076 ... G4300 eG4300 6.020 0.132 6.150 0.094 6.160 0.094 6.030 0.101 5.980 0.114 6.170 0.083 6.250 0.094 7.080 0.131 ... Fe4383 eFe4383 4.840 0.219 5.680 0.149 6.060 0.148 6.110 0.158 4.990 0.194 5.130 0.136 5.000 0.155 5.970 0.218 ... Ca4455 eCa4455 2.160 0.112 2.290 0.079 2.330 0.080 2.340 0.084 2.270 0.094 2.450 0.066 2.530 0.078 2.430 0.108 ... Fe4531 eFe4531 3.820 0.191 4.210 0.157 4.010 0.153 4.130 0.163 3.970 0.163 4.100 0.117 3.700 0.139 4.060 0.193 ... 13 other indices ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... other indices ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... Fe4668 eFe4668 7.580 0.225 7.610 0.226 7.660 0.223 7.680 0.224 7.630 0.225 7.640 0.228 7.670 0.237 8.820 0.181 8.450 0.243 8.360 0.235 8.240 0.174 8.150 0.175 8.220 0.169 7.910 0.298 ... Fe4668 eFe4668 7.080 0.318 7.570 0.234 7.740 0.237 7.630 0.252 8.820 0.267 8.130 0.185 8.120 0.211 7.880 0.299 ... Hβ eHβ 1.460 0.099 1.410 0.098 1.440 0.099 1.420 0.099 1.310 0.098 1.300 0.099 1.340 0.103 1.150 0.079 1.200 0.107 1.210 0.104 1.270 0.075 1.300 0.075 1.190 0.073 1.400 0.131 ... Hβ eHβ 1.540 0.127 1.340 0.104 1.560 0.103 1.330 0.108 1.210 0.104 1.210 0.076 1.230 0.095 1.830 0.123 ... Fe5015 eFe5015 5.860 0.198 5.880 0.198 5.910 0.198 5.950 0.199 5.970 0.206 6.000 0.207 6.090 0.216 6.220 0.176 6.260 0.227 6.190 0.220 6.170 0.165 6.200 0.165 6.230 0.164 6.310 0.293 ... Fe5015 eFe5015 5.520 0.262 5.830 0.200 5.940 0.199 6.330 0.211 6.140 0.233 6.180 0.184 6.330 0.212 5.970 0.302 ... The complete Table 10 is given in electronic form. the majority of the indices α value is very close to 1 and only a zero-point correction is required (see also Puzia et al. (2002), although serious deviations from the one-to-one relation are shown by Ca lines 4227 and 4455 and Fe 4531. 4.5. Estimate of indices measurement errors In order to obtain the errors on each measured index we have used the following procedure. Starting from a given extracted spectrum (aperture or gradient at different galactocentric distances), we have generated a set of 1000 Monte Carlo random modifications, by adding a wave- length dependent Poissonian fluctuation from the corre- sponding spectral noise, σ(λ). Then, for each spectrum, we have estimated the moments of the distributions of the 1000 different realizations of its indices. 5. Results For each galaxy the set of 25 indices obtained for the 7 apertures and the 4 gradients are provided in electronic form with the format shown in Tables 9 and 10 respec- tively. The structure of the above tables is the following: each aperture (or gradient) is described by two rows. In the first raw: col. 1 gives the galaxy identification, col. 2 the number of the aperture, col. 3 is a flag: 0 stands for values of indices, col. 4 and Col. 5 give the radii delimited by the aperture, from col. 6 to 30 individual indices are given. In the second row: col. 1 gives the galaxy identifi- cation, col. 2 the number of the aperture, col. 3 is a flag: 1 stands for error of the indices, col. 4 and Col. 5 give the lu- minosity weighted radius of the aperture and the adopted equivalent radius, from col. 6 to 30 are given the errors of the indices. In electronic form are also available, under request to the authors, the tables of raw indices (before the veloc- ity dispersion correction) as well as the fully calibrated 14 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Fig. 6. Comparison of index measurements of Gonz´alez (1993: open triangles), Trager (1998: full squares), Longhetti et al. (1998: open circles), Beuing et al. (2002: open pentagons) with our data. Solid lines mark the one-to-one relation. Table 11 summarizes the results of the comparison. Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 15 spectra (apertures and gradients) in digital from for each galaxy. In Figures 8-13 we show as examples the trend with radius of the Mg2, Fe(5335) and Hβ indices for the 50 galaxies in the sample (apertures are marked with open squares, gradient with dots). 5.1. Comparison with the literature A significant fraction, about 60%, of galaxies in the present sample has one previous measurement in the Lick -- IDS system but basically restricted within the re/8 region. Line-strength measures obtained both from apertures and gradients outside this area and within the re/8 region, with the present radial mapping, are completely new. The set of line -- strength indices in the literature avail- able for a comparison is quite heterogeneous since indices are measured within different apertures. Furthermore, there are many possible sources of systematic errors from seeing effects to the calibration applied to shift indices to the same spectrophotometric system and to velocity dis- persion correction. Three galaxies, namely NGC 1453, NGC 4552 and NGC 5846 are in the Gonz´alez (1993) sample. Three galax- ies, namely NGC 1553, NGC 6958 and NGC 7135 be- long to the sample observed by Longhetti et al. (1998a). Twenty one galaxies are in the sample provided by Trager et al. (1998) namely NGC 128, NGC 777, NGC 1052, NGC 1209, NGC 1380, NGC 1407, NGC 1426, NGC 1453, NGC 1521, NGC 2749, NGC 2962, NGC 2974, NGC 3489, NGC 3607, NGC 3962, NGC 4552, NGC 4636, NGC 5077, NGC 5328, NGC 7332 and NGC 7377. Eleven galaxies are in the sample recently published by Beuing et al. (2002), namely IC 1459, IC 2006, NGC 1052, NGC 1209, NGC 1407, NGC 1553, NGC 5898, NGC 6868, NGC 6958, NGC 7007 and NGC 7192. A global comparison with the literature is shown shown is Figure 6. In detail: (1) with Longhetti et al. (1998a) the comparison is made with indices computed on the aperture of 2.5′′radius; (2) with Gonz´alez (1993) on re/8 aperture and (3) with Trager et al. (1998) with indices computed within standard apertures; (4) with Beuing et al. (2002) with indices computed on the aperture with ra- dius re/10, taking into account that these authors did not correct Hβ for emission infilling. In Table 11 we present a summary of the comparison with the literature. Both the offset and the dispersion for the various indices in the table are comparable (or bet- ter) of those obtained on the same indices by Puzia et al. (2002) in their spectroscopic study of globular cluster. The comparison of our data with Trager et al. (1998) for each index is in general better than that with other authors and, in particular, with Beuing et al. (2002). The comparison with Trager et al. (1998) shows a zero point shift for Mgb values, our data being larger al- though within the dispersion. A large shift is also shown by G4300 and Ca4227 both also visible in the comparison with Lick stellar indices. Beuing et al. (2002) indices are on consistent range of values, although some systematic effects and zero point offsets are present. While there is a good agreement with the Hβ (without emission correc- tion), Mg1, Mg2 and Fe5335 line -- strength indices, Mgb and Fe5270 show a large zero point differences and dis- persion. Beuing et al. 2002 provided a comparison with Trager et al. (1998), on a partially different sample. They show a basic agreement for the Hβ, Mg1, Mg2 and Mgb (although both a zero point shift and a different slope are visible, e.g. in Hβ and Mg1) while Fe5270, and Fe5335 in- dices show a quite large dispersion and zero point shift as shown in our Figure 6. The well known Mg2 vs. σ relation is plotted in Figure 7. Our Mg2 values computed at the r=1.5′′, compatible with the SLOAN apertures are plotted versus the corre- sponding velocity dispersion values. Adopting the param- eters of the fitting given in Worthey & Collobert (2003) (their Table 1), the dotted line is the least-square fit ob- tained by Bernardi et al. (1998) on the sample of 631 field early-type galaxies while the solid line is the Trager et al. (1998) fit. Bernardi et al. fit is made on the Mg2 index computed using the Lick definition but not transformed to the Lick-IDS system. The long-dashed line shows our least-squares fits to the present data: notice that the value of the Mg index at σ=300 km s−1 of 0.339 is well consis- tent with that of Trager et al. (1998) while the slope for our small set of galaxies is in between those given by the above considered authors. 6. Summary This paper is the first in a series dedicated to the spectro- scopic study of early -- type galaxies showing emission lines in their optical spectrum. It presents the line -- strength in- dex measurements for 50 galaxies residing in cosmological environments of different galaxy richness. The morphol- ogy and the kinematics of the gas component with re- spect to the stellar population are summarized using the available literature in order to characterize each galaxy for subsequent studies. According to current views supported by numerical simulations (see e.g. Colless et al. (1999)) a large fraction of the galaxies in the sample display mor- phological and/or kinematical signatures of past interac- tion/accretion events. For each galaxy, three integrated spectra for different galactocentric distances, center, re/4 and re/2, have been measured. Spectra, in the wavelength range of 3700A< λ < 7250A, have been collated in the Atlas which covers a large set of E and S0 morphological classes. The paper is dedicated to the characterization of the emission galaxy underlying stellar population through the preparation of a data base of their line-strength indices in the Lick-IDS system once corrected for several effects including infilling by emission and velocity dispersion. For each object we extracted 7 luminosity weighted apertures (with radii: 1.5′′, 2.5′′, 10′′, re/10, re/8, re/4 and re/2) corrected for the galaxy ellipticity and 4 gradients (0 16 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Table 11. Comparison with the literature index Ngal offset dispersion units Hb Mg2 Mgb Fe5270 Fe5335 Mg1 Hb Mg2 Mgb Fe5270 Fe5335 Mg1 G4300 Ca4227 Hb Mg2 Mgb Fe5270 Fe5335 Mg1 6 6 6 6 6 6 20 20 20 21 21 20 21 21 11 11 11 11 11 11 G93 + Long98 0.171 -0.006 0.078 0.208 0.160 -0.007 Trager et al. (1998) -0.031 0.004 0.218 0.051 -0.038 -0.002 0.463 0.519 Beuing et al. (2002) -0.059 0.018 0.470 0.339 0.075 0.011 A 0.886 0.014 mag A 0.458 A 0.334 A 0.348 0.011 mag A 0.405 0.016 mag A 0.321 A 0.333 A 0.376 0.013 mag A 0.946 A 0.757 A 0.271 0.025 mag A 0.628 A 0.603 A 0.455 0.018 mag Offsets and dispersions of the residuals between our data and the literature. Dispersions are 1 σ scatter of the residuals. Beuing et al. (2002) do not compute G4300 and Ca4227 indices. For G4300 and Ca4227 indices Gonz´alez (1993) and Longhetti et al. (1998) used a (slightly) different definition than Trager et al. (1998) and consequently comparisons are not reported in the table. The comparison for Hβ index between our data and those of Trager et al. (1998) and Beuing et al. (2002) are made using uncorrected data. in digital form. RR acknowledges the partial support of the ASI (contract I/R/037/01). WWZ acknowledges the support of the Austrian Science Fund (project P14783) and of the Bundesministerium fur Bildung, Wissenschaft und Kultur. References Barnes, J. 1996 in Galaxies: Interactions and Induced Star Formation, Saas-Fee Advanced Course 26. Lecture Notes 1996. Swiss Society for Astrophysics and Astronomy, XIV, Springer-Verlag Berlin/Heidelberg, p. 275 Bergeron, J., Durret, F., Boksenberg, A. 1983, A&A, 127, 322 Bernardi, M., Renzini, A., Da Costa, L.N. et al. 1998, ApJ, 508, L43 Bertin, G., Bertola, F., Buson, L. et al. 1994, A&A, 292, 381 Bertola, F., Galletta, G., Zeilinger, W.W. 1985, ApJ, 292, L51 Bertola, F., Buson, L.M., Zeilinger, W.W. 1992, ApJ, 401, L79 Bertola, F., Galletta, G., Zeilinger, W.W. 1992, A&A, 254, 89 Bettoni, D., Galletta, G. 1997, A&AS, 124, 61 Bettoni, D., Galletta, G. Garcia-Burillo, S., Rodriguez-Franco, A. 2001, A&A, 374, 421 Beuing, J., Bender, R., Mendes de Oliveira, C., Thomas, D., Maraston, C. 2002, A&A, 395, 431. Fig. 7. Mg2 versus σ relation. Our Mg2 line-strength indices measured within the SLOAN aperture (r=1.5′′) are plotted after the transformation to the Lick-IDS sys- tem. The solid dashed line is the least-square fit obtained Trager et al. (1998). The the dotted and long-dashed lines represent the least-square fit obtained by Bernardi et al. (1998) for the field sample of 631 galaxies and our fit (value at σ300 km s−1 = 0.339, slope of relation = 0.214) to the present data respectively. ≤ r ≤re/16, re/16 ≤ r ≤re/8, re/8 ≤ r ≤re/4 and re/4 ≤ r ≤re/2). For each aperture and gradient we measured 25 line -- strength indices: 21 of the set defined by the Lick-IDS "standard" system (Trager et al. 1998) and 4 introduced by Worthey & Ottaviani (1997). Line-strength indices, in particular those used to build the classic Hβ-< M gF e > plane, have been compared with literature. A direct comparison was made with Gonz´alez (1993), Longhetti et al. (1998a), Beuing et al. (2002) but in particular with the larger data sets of Trager et al. (1998) showing the reliability of our measures. In forthcoming papers we plan (1) to model this in- formation for investigating the ages and metallicities of the bulk of the stellar populations in these galaxies; (2) to extend the study calibrating and modeling other "blue indices" outside the Lick/IDS system namely ∆(4000A) (Hamilton 1985) , H+K(CaII) and Hδ/FeI (Rose 1984; Rose 1985) correlated with the age of the last starburst event in each galaxy; (3) to complete the study of the emission line component on a finer spatial grid. Acknowledgements. The authors would like to thank the ref- eree, Guy Worthey, whose comments and suggestions signif- icantly contributed to improve the paper. We are deeply in- debted with Enrico V. Held for his invaluable help. We thank Daniela Bettoni, Nicola Caon and Francois Simien for mak- ing their velocity dispersion measurements available to us Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 17 Binette, L., Magris, C.G., Stasinska, G., Bruzual, A.G. 1994, Longhetti, M., Bressan, A., Chiosi, C., Rampazzo, R. 1999, A&A, 292, 13 A&A, 345, 419 Birkinshaw, M., Davies, R.L. 1985, ApJ, 291, 32 Blanton, E. L.; Sarazin, C. L.; Irwin, J. A. 2001, ApJ, 552, 106 Bregman, J.N., Hogg, D.E., Roberts, M.S. 1992, ApJ, 397, 484. Buson, L.M., Sadler, E.M., Zeilinger, W.W. et al., 1993, A&A, Longhetti, M., Bressan, A., Chiosi, C., Rampazzo, R. 2000, A&A, 353, 917 Longo, G., Zaggia, S.R., Busarello, G. et al. 1994, A&AS, 105, 433 280, 409 Macchetto, F., Pastoriza, M., Caon, N. et al. 1996, A&AS, 120, Buzzoni, A., Gariboldi, G., Mantegazza, L. 1992, AJ, 103, 1814 Buzzoni, A., Mantegazza, L. Gariboldi, G. 1994, AJ, 107, 513 Caldwell, N., Kirshner R.P., Richstone, D.O. 1986, ApJ, 305, 463 Malin, D.F., Carter, D. 1983, ApJ, 274, 534 Maoz, D., Koratkar, A., Shields, J.C. Ho., L.C., et al. 1998, 136 AJ, 116, 55 Caon N., Macchetto, D., Pastoriza, M. 2000, ApJS, 127, 39 Capaccioli,M., Piotto, G., Rampazzo, R. 1988, AJ, 96, 487 Carollo, M., Danziger, I.J., Buson, L.M. 1993, MNRAS, 265, 553 Carollo, M., Danziger, I.J. 1994, MNRAS, 270, 523 Carrasco, L., Buzzoni, A., Salsa, M., Recillas-Cruz, E. 1995, in Fresh Views on Elliptical Galaxies, ed.s A. Buzzoni, A. Renzini, A. Serrano, ASP Conf. Series 86, 175 Cinzano, P., van der Marel, R.P. 1994, MNRAS, 270, 325 Colbert, J.W., Mulchaey, J.S., Zabludoff, A.I. 2001, AJ, 121, 808 Colina, L., Sparks, W.B., Macchetto, F. 1991, ApJ, 370, 102 Colles, M., Burstein, D., Davies, R., McMahan, R.K., saglia, Mehlert, D., Thomas, D., Saglia, R.P., Bender, R.,,Wegner, G. 2003, A&A, 407, 423 Michard, R., Marchal, J. 1994, A&AS, 105, 481 Morganti, R. Oosterloo, T., Tsvetanov, Z. 1998, AJ, 115, 915 Osterbrock, D., 1989, in Astrophysics of Planetary Nebulae and Active Galactic Nuclei, University Science Books Pizzella, A., Amico,P., Bertola,F. et al. 1997, A&A, 323, 349 Plana, H., Boulesteix, J. 1996, A&A, 307, 391 Plana, H., Boulesteix, J. Amram, Ph., Carignan, C., Mendes de Oliveira, C. 1998, A&AS, 128, 75 Phillips, M., Jenkins, C., Dopita, M., Sadler, E.M., Binette, L. 1986, AJ, 91, 1062 Phillips, A.C., Illingworth, G.D., MacKnety, J.W. et al. 1996, R.P., Wegner, G. 1999, MNRAS, 303, 813 AJ, 111, 1566 De Vaucouleurs, G., de Vaucouleurs, A., Corwin, H.G. Jr. et al. 1991 Third Reference Catalogue of Bright Galaxies, Springer-Verlag, New York Djorgovsky, S., Davis, M. 1987, ApJ, 313, 59 D'Onofrio, M., Zaggia S.R., Longo, G., Caon, N., Capaccioli, M. 1995, A&A, 296, 319 Prugniel, Ph., Simien, F. 1997, A&AS, 122, 521 Prugniel, Ph., Simien, F. 1998, A&AS, 131, 287 Prugniel, Ph., Simien, F. 2000, A&AS, 145, 263 Puzia, T.H., Saglia, R.P., Kissler-Patig, M., Maraston, C., Greggio, L. et al. 2002, A&A, 395, 45 Quillen, A.C., Bower, G.A., Stritzinger, M. 2000, ApJS, 128, D'Onofrio, M., Capaccioli, M., Merluzzi, P., Zaggia, S., 85 Boulesteix, J. 1999, A&AS, 134, 437 Raimann, D., Storchi-Bergmann, T., Bica, E., Alloin, D. 2001, Emsellem, E., Arsenault, R. 1997, A&A, 318, L39 Franx, M., Illingworth, G.D., Heckman, T. 1989, ApJ, 344, 613 Franx, M., Illingworth, G.D. 1988, ApJ, 327, L55 Fried, J.W., Illingworth, G.D. 1994, AJ, 107, 992 Gabel, J.R., Bruhweiler, F.C., Crenshaw, D,M., Kraemer, S.B., MNRAS, 324, 1087 Rampazzo, R. 1988, A&A, 204, 81 Rampazzo, R., Plana, H., Longhetti, M. et al. 2003, MNRAS, 343, 819 Roberts, M.S., Hogg, D.E., Bregman, J.N., Forman, W.R., Miskey, C.L. 2000, ApJ, 532, 883 Jones, C. 1991, ApJS, 75, 751 Gonz´alez, J.J. 1993, Ph.D. thesis , Univ. California, Santa Cruz. Goudfrooij, P. 1994, Ph.D. thesis, University of Amsterdam, The Netherlands Goudfrooij, P. 1998, in Star Formation in Early -- Type Galaxies, ASP Conference Series 163, ed.s P. Carral and J. Cepa, 55 Goudfrooij, P., Trinchieri, G. 1998, A&A, 330, 123 Hamilton, D. 1985, ApJ, 297, 371 Heckman, T.M. 1980, A&A, 87, 152 Kim, D.-W. 1989, ApJ, 346, 653 Koprolin, W., Zeilinger, W.W. 2000, A&A, 145, 71 Kormendy, J. 1984, ApJ, 286, 116 Kuntschner, H. 2000, MNRAS, 315, 184 Kuntschner, H., Smith, R.J., Colless, M. et al. 2002, MNRAS, 337, 172 Rose J.A. 1984, AJ, 89, 1238 Rose J.A. 1985, AJ, 90, 1927 Saglia, R.P., Bertin, G., Bertola, F. et al. 1993, ApJ, 403, 567 Sandage, A.R., Tammann, G. 1987, A Revised Shapley Ames Catalogue of Bright Galaxies, Carnegie, Washington (RSA) Saraiva, M.F., Ferrari, F., Pastoriza, M.G. 1999, A&A, 350, 339 Schweizer, F., van Gorkom, G.H., Seitzer, P. 1989, ApJ, 338, 770 Schweizer, F. 1996 in Galaxies: Interactions and Induced Star Formation, Saas-Fee Advanced Course 26. Lecture Notes 1996. Swiss Society for Astrophysics and Astronomy, XIV, Springer-Verlag Berlin/Heidelberg, p. 105 Sharples, R.M., Carter, D., Hawarden, T.G., Longmore A.J. 1983, MNRAS, 202, 37 Jarvis, B. 1987, AJ, 94, 30 Jedrzejewski, R., Schechter, P.L. 1989, AJ, 98, 147 Lauberts, A., Valentijn E.A. 1989, The Surface Photometry Catalogue of the ESO-Uppsala Galaxies, ESO. Simien, F., Prugniel, Ph. 1997a, A&AS, 126, 15 Simien, F., Prugniel, Ph. 1997b, A&AS, 126, 519 Thomas, D., Maraston, C., Bender, R. 2003, MNRAS, 339, 897 Trager, S.C., Worthey, G., Faber, S.M., Burstein, D., Gozalez Leonardi, A.J, Rose, J.A. 1996, AJ, 111, 182 Longhetti, M., Rampazzo, R., Bressan, A., Chiosi, C. 1998a, J.J. 1998, ApJS, 116, 1 Trager, S.C., Faber, S.M., Worthey, G., Gozalez J.J. 2000, AJ, A&A, 130, 251 119, 164 Longhetti, M., Rampazzo, R., Bressan, A., Chiosi, C. 1998b, Trinchieri, G., Noris, L., di Serego Alighieri, S. 1997, A&A, A&A, 130, 267 326, 565 18 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Tully, R.B. 1988, Nearby Galaxy Catalogue, Cambridge University Press Ulrich-Demoulin M-H. et al. 1984, ApJ, 285, 13 van Driel, W., Arnaboldi, M., Combes, F., Sparke, L.S. 2000, A&AS, 141, 385 van Gorkom , J.H., Knapp, G.R., Raimond, E. et al. 1986, AJ, 91, 791 Varnas, S.R., Bertola, F., Galletta, G., Freeman, K.C., Carter, D. 1987, ApJ, 313, 69 Vazdekis A. 1999, ApJ, 513, 224 V`eron-Cetty, M.-P, V`eron, P. 2001, A&A, 374, 92 Voit, G.M. 1991, ApJ 377, 158 White III, R., Sarazin, C.L. 1991, ApJ, 367, 476 Wilkinson, A., Sharples, R.M., Fosbury, R.A.E., Wallace, P.T. 1986, MNRAS, 218, 297 Worthey, G. 1992, Ph.D. Thesis, University of California, Santa Cruz Worthey, G., Faber, S.M., Gonz´alez, J.J., Burstein, D. 1994, ApJS, 94, 687 Worthey, G., Ottaviani, D.L. 1997, ApJS, 111, 377 Worthey, G., Collobert, M. 2003, ApJ, 586, 17 Wrobel, J.M., Heeschen, D.S. 1984, ApJ, 287, 41 Zeilinger, W.W., Pizzella, A., Amico, P., Bertin, G. et al., 1996, A&AS, 120, 257 Appendix A: Relevant notes on individual galaxies from the literature We report in this section some studies relevant to the present investigation performed in the recent literature with particular attention to the properties of the ionized gas with respect to the bulk of the stellar component. NGC 128 Emsellem & Arsenault (1997) present a study of the gas (and dust) disk tilted at an angle of 26◦ with respect to the major axis of the galaxy. The stellar and gas velocity fields show that the angular momentum vectors of the stellar and gaseous components are reversed, suggesting that the gas component orbits suffer the pres- ence of a tumbling bar, possibly triggered by the interac- tion of NGC 128 with the nearby companion NGC 127. The gas extends at least up to 6′′ from galaxy center, and in the inner parts line ratios are typical of LINER and consistent with the gas being ionized by post-AGB stars. D'Onofrio et al. (1999) evaluate the central gas mass is ≈ 2.7 104 M⊙. The dust does not have the same distribution as the gas but is largely confined to the region of interac- tion between NGC 128 and NGC 127. They calculated a dust mass of ≈ 6×106 M⊙ for NGC 128. NGC 777 The kinematics and the photometry of this galaxy were obtained by Jedrzejewski & Schechter (1989). Both the P.A. and the ellipticity profile appear nearly constant at about 149◦ and 0.18 respectively. Both the kinematics along the major and minor axes have been in- vestigated. A rotation of about 50 km s−1 is measured along the major axis while no apparent rotation is de- tected along the minor one. NGC 1052 The galaxy is known as a prototypical LINER (Heckman 1980). Plana et al. (1998), perform- ing Fabry-Perot observations, succeeded in disentangling two gas components both kinematically decoupled from the stellar component. Both, in fact, have their apparent major axis nearly perpendicular to the stellar one. The ionized gas of the main component was detected up to 30′′ from the center, while the second one extends up to 15′′. The main component shows rotation with an appar- ent major axis of 45◦± 4◦ similar to that of the HI emis- sion detected by van Gorkom et al. (1986) with which it shares also similar kinematical characteristics. The veloc- ity field of both components presents shapes and velocity dispersion in agreement with models of inner disks found in elliptical galaxies. Recently Gabel et al. (2000) imaged the central part of the galaxy using WFPC2/HST with an Hα filter showing that a filamentary nebular emission extends about 1′′ around a compact nucleus with a more diffuse halo extending to further distances. At the posi- tion angle of ≈ 235◦ there is a narrow filament of Hα emission. A radio jet/lobe at a position angle of ≈ 275◦ has been evidenced by Wrobel & Heeschen (1984). The emission line region is much more extended, as discussed above. Gabel et al. (2000) examined whether or not the ionizing continuum flux is sufficient to power the above extended emission line region. They conclude that a pure- central source photo-ionization model with the simplest non-thermal continuum (a simple power law) reproduces the emission line flux in the inner region of NGC 1052. Other processes, such as shocks or photoionization by stars, are not required to produce the observed emission. However, the contribution of these latter mechanisms can- not be ruled out in the extended nebular emission region. Recently Raiman et al. (2001) have analyzed the spectra of NGC 1052 and IC 1459, classified as LINERS, at sev- eral galactocentric distances from the nucleus. They found that these objects have both the nucleus redder than the surroundings and nuclear absorption lines stronger than outside the nucleus similarly to normal galaxies. On the other side the spectral synthesis of NGC 1052 and IC 1459 indicate that they have only a ≈ 10 - 20 % larger contribution of the 1-Gyr component at the nucleus with respect to normal galaxies which are dominated by the old metal rich component whose contribution is decreas- ing outwards. The above authors exclude the presence of young massive stars found by Maoz et al. (1998). NGC 1209 The galaxy is part of the group dominated by NGC 1199. The surface photometry and the geometri- cal study, performed by Capaccioli et al. (1988), extends up to µB ≈ 28 mag arcsec−2. The ellipticity grows from 0.22 to 0.57 with basically no twisting (< P.A > = 51◦ ±2◦ up to 126′′) suggesting that the galaxy is an S0. NGC 1380 D'Onofrio et al. (1995) suggest the presence of a dusty nucleus. The stellar component has a strong gradient in both rotation and velocity dispersion curves and the disk dominates outside 20′′. Kuntschner (2000), studying stellar populations of early-type galaxies in the Fornax cluster, found that this bright lenticular has prop- erties similar to those of ellipticals suggesting that they experienced similar star formation histories. NGC 1380 ex- hibits an overabundance in magnesium compared to iron, similarly to most ellipticals in the Fornax cluster. Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 19 NGC 1389 This galaxy belongs to the Fornax cluster. Phillips et al. (1996) examined the nucleus of the galaxy using the HST Planetary Camera, finding no evidence for an unresolved central point source. The image shows a smooth light distribution sharply peaked at the center and isophote twisting within 5′′. NGC 1407 The (B-V) image of this galaxy, which is a radio source, reveals a circumnuclear ring slightly redder than the nucleus (Goudfrooij 1994). The stellar kinematics were studied along several axes by Longo et al. (1994). Franx et al. (1989) found significant rotation along the minor axis. NGC 1426 The study performed by Capaccioli et al. (1988) extends up to µB ≈ 28 mag arcsec−2 and indicates that the galaxy is an S0 with basically no isophotal twist- ing (< P.A > = 105◦ ±1◦ up to 100′′). Quillen (2000) have observed the galaxy with HST NICMOS. The cen- tral 10′′ show very regular isophotes with no twisting or deviations from elliptical shapes. The core has a power law profile and no dust features are detected. The extended ro- tation curve and velocity dispersion profiles, obtained by Simien & Prugniel (1997a), do not show any peculiarities. NGC 1453 Pizzella et al. (1997) found that this E2 galaxy has a twisting of 10◦. The Hα image reveals the presence of an ionized disk misaligned with respect to the stellar isophotes by ≈ 56◦ suggesting an intrinsic triaxial shape. NGC 1521 The surface photometry and the geometri- cal study performed by Capaccioli et al. ( 1988) extends up to µB ≈ 27-28 mag arcsec−2 at 3.8 re. The galaxy shows a peculiar light distribution with a change in slope along the major axis and a significant twisting (≈=20◦ up to 126′′). The extended rotation curve and velocity dis- persion profiles, obtained by Simien & Prugniel (1997a), do not show peculiar features. NGC 1553 The galaxy belongs to the shell galaxies sample of of Malin & Carter (1983). The galaxy kinematics is typical of an early S0s and shows two maxima in the ro- tations curve (Kormendy 1984; Rampazzo 1988; Longhetti et al. 1998b). No rotation is found along the minor axis. The Hα narrow band image of this galaxy (Trinchieri et al. 1997) shows a strong nuclear peak and a bar-like fea- ture ≈ in the North-South direction that ends in spiral structure at 8′′ from the nucleus. Blanton et al. (2001) pro- posed using Chandra data that the center of NGC 1553 is probably an obscured AGN while, the X-ray diffuse emis- sion exhibits significant substructure with a spiral feature passing through the center of the galaxy. Longhetti et al. (1998a, 1999, 2000) studied the stellar population of this galaxy suggesting that the age of a secondary burst is old, probably associated to the shell formation. Rampazzo et al. (2003) measured the velocity field of the gas component using Fabry-Perot data. In the central region of NGC 1553 the ionized gas is co-rotating with the stellar component. NGC 1947 The galaxy is considered a minor-axis dust- lane elliptical (Bertola et al. 1992b). The stellar compo- nent of this galaxy is rotating around the minor axis, per- pendicularly to the gas rotation axis. The gas component forms a warped disk whose external origin is suggested by Bertola et al. (1992b). NGC 2749 The galaxy, studied by Jedrzejewski & Schechter (1989), shows strong rotation (≥ 100 km s−1) along both the major and minor axes. A measure of the gas rotation curve was attempted using the λ 5007 A line. They suggested that the gas is not rotating, with an upper limit of the order of one-half the stellar rotation velocity on either axis. NGC 2911 Known also as Arp 232 the galaxy is clas- sified as a LINER in the V`eron-Cetty & V`eron (2001) catalog. Michard & Marchal (1994) suggest that this is a disk dominated S0 with a significant dust component. NGC 2962 The extended rotation curve and velocity dispersion profiles, obtained by Prugniel & Simien (2000), do not show peculiarities. NGC 2974 This E4 galaxy imaged in Hα reveals the presence of an ionized disk misaligned with respect to the stellar isophotes by ≈ 20◦ (Pizzella et al. 1997; Ulrich- Demoulin et al. 1984; Goudfrooij 1994). The galaxy has an HI disk (Kim 1989) with the same rotation axis and velocity as the inner ionized one. Plana et al. (1998) sug- gest that this object is a good candidate for an internal origin of the ionized gas. Bregman et al. (1992) present evidence of a spiral arm structure and Cinzano & van der Marel (1994) could not discard the hypothesis that NGC 2974 is a Sa galaxy with an unusually low disk-to-bulge ratio. NGC 3136 Using an Hα+[NII] image Goudfrooij (1994) detected an extended emission with a maximum at the nucleus and peculiar arm-like feature extending out to ≈ 55′′ from the center. Dust absorption is found to be as- sociated with the ionized gas. Koprolin & Zeilinger (2000) suggest that a counterrotating disk with a dimension of 2′′ is located 4′′ from the galaxy center. NGC 3258 Koprolin & Zeilinger 2000 measured a very low rotation velocity of 39±10 km s−1 for this galaxy. NGC 3268 Koprolin & Zeilinger (2000) found that the galaxy has an asymmetric rotation curve with respect to the nucleus, probably due to the presence of a dust-lane. NGC 3489 Gas and stars show a fast rotation along the major axis of the gas distribution which roughly coincides with the major axis of the stellar isophotes. The gas shows rotation along the minor axis while no stellar rotation is measured. There is evidence for a distinct nuclear stellar component (within r ≈ 3′′) (Caon et al. 2000). NGC 3557 The color map reveals a possible ring of dust near the center of the galaxy (Colbert et al. 2001). The galaxy is a double tail radio source with a central knot and a jet (Birkinshaw & Davies 1985). Goudfrooij (1994), using an Hα+[NII] image, shows that the outer isophotes of the line emission twist gradually toward the apparent major axis of the galaxy. NGC 3607 Caon et al. (2000) observe stellar kinemat- ics along the major axis which is also the major axis of the gas distribution. The gas rotation curve has a steeper gradient and a larger amplitude than the stellar one. 20 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. NGC 3962 Birkinshaw & Davies (1985) revealed a ra- dio source in the center of the galaxy. The morphology and the kinematics of the ionized gas confirm the pres- ence of two distinct subsystems: an inner gaseous disk and an arc-like structure. The inner gaseous disk shows regular kinematics with a major axis near P.A.= 70◦ and inclination of about 45◦ (Zeilinger et al. 1996). NGC 4552 The extended kinematics of this galaxy has been recently obtained by Simien & Prugniel (1997b) who measure a very low maximum rotation velocity of 17±10 km s−1. NGC 4552 is a member of the Malin & Carter (1983) supplementary list of galaxies showing shells (they report "two or three shells and jet") NGC 4636 Caon et al. (2000) observed the galaxies along three axes and found that the gas has very irregular velocity curves. Zeilinger et al. (1996) suggested that the gas could suffer for turbulent motions due to material not yet settled. NGC 5077 Caon et al. (2000) found that the galaxy exhibits a gaseous disk with major axis roughly orthogonal to the galaxy photometric major axis. The gas isophotes show a twisting and a warp (Pizzella et al. 1997). The gas has a symmetric rotation curve with an amplitude of 270 km s−1 at r=13′′. Along this axis the stellar rotation is slow. Along the axis at P.A.=10◦, the stellar velocity curve shows a counterrotation in the core region (r<5′′) with a corresponding peak in the stellar velocity dispersion. NGC 5266 This galaxy has a dust lane along the ap- parent minor axis of the elliptical stellar body. The kine- matics of NGC 5266 has been extensively studied by Varnas et al. (1987) revealing a cylindrical rotation of the stellar component (Vmax=212 ± 7 km s−1) about the short axis and smaller rotation (V=43 ± 16 km s−1) about the long axis. The stellar velocity rotation is 210 ± 6 km s−1 and decreases with radius to 100 km s−1 at r ≈ 20′′. The gas associated with the dust rotates about the major axis of the galaxy with a velocity of 260 ± 10 km s−1. In the warp the gas motion are prograde with respect to the major axis stellar rotation. HI radio obser- vations reveal the presence of a large amount of cold gas probably distributed in a rotating disk. NGC 5363 The galaxy has a warped dust lane con- fined to the central part along its apparent minor axis. Differently from NGC 5128, the gas motions in the warp are found to be retrograde with respect to the stellar body. Bertola et al. (1985) suggest that is an indication that the warp is a transient feature and of the external origin of the gas and dust system. NGC 5846 Ulrich-Demoulin et al. (1984) have studied the ionized gas component in this galaxy. Several studies suggests the presence of dust (see Goudfrooij 1998) in the galaxy. Caon et al. (2000) found that the gas shows an irregular velocity profile, while stars have very slow rota- tion. NGC 5898 Caon et al. (2000) analyzed the stellar and the gas kinematics up to 45′′ showing the existence of a stellar core of 5′′ in radius, aligned with the major axis, which counterrotates with respect to the outer stellar body. The ionized gas counterrotates with respect to the inner stellar core and co-rotates with respect to outer stel- lar body. At the same time, the gas counterrotates along the minor axis, indicating that the angular momentum vectors of the stars and of the gas are misaligned, but not anti-parallel. A moderate quantity of dust is also present. NGC 6721 Bertin et al. (1994 ) obtained extended stel- lar kinematics (to 0.8 re) for this object finding a sizeable rotational velocity ≈ 120 km s−1. NGC 6868 The Fabry-Perot study of Plana et al. (1998) shows that the line -- of -- sight velocity field of the ionized gas component has a velocity amplitude of ± 150 km s−1. Caon et al. (2000) show that along the axes at P.A.=30◦ and 70◦ the gas and stars have similar kinemat- ical properties, but along P.A.=120◦ the gas counterro- tates with respect to the stellar component. Zeilinger et al. (1996) noticed the presence of an additional inner gas com- ponent which suggested could be due to the superposition of two unresolved counterrotating components, one dom- inating the inner region, the other dominating the outer parts. Also stars show a kinematically -- decoupled coun- terrotating core. The stellar velocity dispersion decreases towards the galaxy center. NGC 6958 This galaxy belongs to the list of Malin & Carter (1983) of southern shell early-type. Saraiva et al. (1999) detected isophotal twisting of about 100◦ (≈ 70◦ in the inner 5′′) but no particular signature of interaction in the isophote shape which is elliptical. They conclude that if the galaxy suffered interaction, the companion galaxy has probably already merged. NGC 7007 Pizzella et al. (1997) found that the disk is misaligned by about 30◦ with respect to the stellar isophotes with an inclination of 57◦. The ionized gas disk counterrotates with respect to the stellar body and a bow- shape dust lane is also visible on the eastern side (Bettoni et al. 2001). NGC 7079 A counterrotating disk-like structure of ion- ized gas within 20′′ from the center has been detected by Bettoni & Galletta (1997). The stellar body kinematics is typical of an undisturbed disk. Cool gas (CO) has been detected by Bettoni et al. (2001). The cool gas component shares the same kinematics of the ionized gas. NGC 7097 Caldwell et al. (1986) found that the gas and stellar components counterrotate. Zeilinger et al. (1996) show that the rotation curve of the gaseous disk have a steep central gradient with a discontinuity in the central part which may be related to the counterrotating stellar component. NGC 7135 The galaxy belongs to the list of shell galaxies in the southern hemisphere compiled by Malin & Carter (1983). Longhetti et al. (1998a, 1998b, 1999, 2000) studied the inner kinematics and the stellar population of this galaxy. Rampazzo et al. (2003) show that the gas corotates with the stellar body. NGC 7192 Carollo & Danziger (1994) showed that the innermost 8′′ region counterrotates with respect to the galaxy body at greater radii. The authors report that in correspondence with the kinematically decoupled core an Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 21 radio knots in IC 5063. It has been estimated (Morganti et al. 1998) that the energy flux in the radio plasma is an order of magnitude smaller than the energy flux emit- ted in the optical emission lines. The shocks associated with the jet-ISM interaction are, therefore, unlikely to ac- count for the overall ionization, and the NLR must be, at least partly, photo-ionized by the nucleus, unless the lobe plasma contains a significant thermal component. enhancement in the Mg2 index is observed, while iron lines are only weakly enhanced with respect to measurements at greater radii. The surface photometry shows that the galaxy has a very regular, round structure. NGC 7332 Plana & Boulesteix (1996), using a Fabry- Perot (CIGALE), found two gas components (see also NGC 1052). The velocity field is consistent with two coun- terrotating emission systems. IC 1459 This giant elliptical has a massive counter- rotating stellar core (M≈1010 M⊙; Franx & Illingworth (1988) which hosts a compact radio source. The galaxy is also crossed by a disk of ionized gas (see §2.1), whose emis- sion is detected out to 35′′ and rotates in the same direc- tion as the outer stellar component but at a higher speed (350 km s−1). Therefore counterrotation in this galaxy seems confined to the inner core and affects only stars. Bettoni et al. (2001) detect 12CO(J=2-1) emission decou- pled both from the ionized gas and the counterrotating stellar core, since the velocity centroid of the CO emis- sion is redshifted by about 100 km s−1 with respect to the galaxy systemic velocity. Raimann et al. (2001) presented spectral syntheses at several galacto-centric distances (see notes for NGC 1052). IC 2006 A faint emission of ionized gas, characterized by a velocity gradient which is smaller and inverted with respect to that of stars, is detected within 25′′ of the nu- cleus. The counterrotating ionized gas disk is highly tur- bulent with a measured velocity dispersion of 190 km s−1 (Schweizer et al. 1989). Kuntschner (2000) found that IC 2006 has stronger metal-line absorption than what would be expected from the mean index-σ0 relation for ellipti- cals, although from their data it is not clear if the galaxy is too metal-rich or whether the central velocity dispersion is lower than for other ellipticals of the same mass. IC 3370 The galaxy is a box shaped elliptical, with a prominent dust-lane in the inner region, showing evidence of cylindrical rotation and X-shaped isophotes. Van Driel et al. (2000) postulated that it is a candidate polar-ring galaxy. IC 4296 The galaxy kinematics have been studied by Franx et al. (1989) and more recently by Saglia et al. (1993) out to 0.8re confirming the counterrotating core, detected by previous authors. Saglia et al. suggest the presence of a diffuse dark matter halo. IC 5063 Long -- slit spectroscopy for the radio galaxy IC 5063 has uncovered a clear rotation pattern for the gas close to the nucleus and a flat rotation curve further out, to at least 19 kpc (Bergeron et al. 1983). The velocity dif- ference between the two flat parts of the rotation curve on both sides of the nucleus is ∆V = 475 km s−1. Colina et al. (1991) report a very high excitation emission line spec- trum for this early-type galaxy which hosts a Seyfert 2 nu- cleus that emits particularly strongly at radio wavelength. The high excitation lines are detected within 1-1.5′′ of both sides of the nucleus, which is approximately the dis- tance between the radio core and both of the lobes. These lines indicate the presence of a powerful and hard ioniz- ing continuum in the general area of the nucleus and the 22 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Fig. 8. Fully corrected Mg2 line-strength index as function of the luminosity weighted radius normalized to the galaxy equivalent radius Re. Apertures are indicated with open squares, while gradients are indicated with full dots. Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 23 Fig. 9. As in Figure 8. 24 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Fig. 10. Fully corrected Hβ line-strength index as function of the luminosity weighted radius normalized to the galaxy equivalent radius Re. Apertures are indicated with open squares, while gradients are indicated with full dots. Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 25 Fig. 11. As in Figure 10. 26 Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. Fig. 12. Fully corrected Fe5335 line-strength index as function of the luminosity weighted radius normalized to the galaxy equivalent radius Re. Apertures are indicated with open squares, while gradients are indicated with full dots. Rampazzo et al.: Nearby early -- type galaxies with ionized gas.I. 27 Fig. 13. As in Figure 12. 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 4000 5000 6000 7000 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 4000 5000 6000 7000 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 1.5 1 0.5 4000 5000 6000 7000
astro-ph/0702484
3
0702
2007-11-11T14:51:50
The LambdaCDM model on the lead -- a Bayesian cosmological models comparison
[ "astro-ph" ]
Recent astronomical observations indicate that our Universe is undergoing a period of an accelerated expansion. While there are many cosmological models, which explain this phenomenon, the main question remains which is the best one in the light of available data. We consider ten cosmological models of the accelerating Universe and select the best one using the Bayesian model comparison method. We demonstrate that the LambdaCDM model is most favored by the Bayesian statistical analysis of the SNIa, CMB, BAO and H(z) data.
astro-ph
astro-ph
The ΛCDM model on the lead -- a Bayesian cosmological models comparison Aleksandra Kurek Astronomical Observatory, Jagiellonian University, Orla 171, 30-244 Krak´ow, Poland Marek Szyd lowski Astronomical Observatory, Jagiellonian University, Orla 171, 30-244 Krak´ow, Poland Complex Systems Research Centre, Jagiellonian University, Reymonta 4, 30-059 Krak´ow, Poland ABSTRACT Recent astronomical observations indicate that our Universe is undergoing a period of an accelerated expansion. While there are many cosmological models, which explain this phenomenon, the main question remains which is the best one in the light of available data. We consider ten cosmological models of the accel- erating Universe and select the best one using the Bayesian model comparison method. We demonstrate that the ΛCDM model is most favored by the Bayesian statistical analysis of the SNIa, CMB, BAO and H(z) data. 1. Introduction Recent observations of type Ia a supernovae (SNIa) provide the main evidence that the current Universe is in an accelerating phase of expansion (Riess et al. 1998; Perlmutter et al. 1999). There are many different cosmological models used in explanation of an accelerating phase of evolution of the current Universe. They can be divided into two groups of models according to 'philosophical' assumptions on a cause of the accelerated expansion. In the first type of explanation the conception of mysterious dark energy of an unknown form is used, while in the second one it is postulated some modification of the Friedmann equation. Here we choose five models which belong to the former group as well as five ones which belong to the latter one. All the chosen models are assumed to be spatially flat. If we assume the Friedmann-Robertson-Walker (FRW) model in which effects of non- homogeneities are neglected, than acceleration can be driven by a dark energy component X (matter fluid violating the strong energy condition). This kind of energy represents -- 2 -- roughly 70% of the matter content of the present Universe. The model with the cosmological constant (the ΛCDM model) has the equation of state for dark energy as follows: pX = −ρX (Weinberg 1989). The model with phantom dark energy has pX = wXρX , where wX (< −1) is a negative constant (Caldwell 2002; Dabrowski et al. 2003). The next one is the model with a dynamical coefficient of the equation of state, parameterized by the scale factor a: w(a) = w0 +w1(1−a) (Chevallier & Polarski 2001; Linder 2003). The other simple approach is to represent dark energy in the form of a minimally coupled scalar field φ with the potential V (φ). In cosmology the quintessence idea is important in understanding a role of the scalar field in the current Universe. We consider the power-law parameterized quintessence model (Peebles & Ratra 1988; Ratra & Peebles 1988). In this case density of dark energy changes with the scale factor as ρX = ρX0a−3(1+ ¯wX (a)), where ¯wX(a) is the mean of a coefficient of the equation of state in the logarithmic scale factor ¯wX(a) = R wX(a)d ln a R d(ln a) and has the following form ¯wX = w0aα (Rahvar & Movahed 2007). The first group is com- pleted with the model with the generalized Chaplygin gas, where pX = − A (here A > 0 and α = const). We gathered above models together with their Hubble functions (with the assumption that the Universe is spatially flat) in Table 1. ρα X As we have written before we also consider five models offering the explanation of cur- rent acceleration of the Universe in an alternative way to dark energy. The brane models has postulated that the observer is embedded on the brane in a larger space in which grav- ity can propagate: the Dvali-Gabadadze-Porrati model (DGP) (Dvali et al. 2000), Sahni- Shtanov brane 1 model (Shtanov 2000). The Cardassian model, in which the Universe is flat, is matter dominated and accelerating as a consequence of the modification of the Friedmann first integral as follows 3H 2 = ρ + Bρn, where B is a constant and the energy Table 1. The Hubble function for cosmological models with dark energy case model H 2(z)relation 1 ΛCDM model H 2 = H 2 0 {Ωm,0(1 + z)3 + (1 − Ωm,0)} 2 model with generalized Chaplygin gas H 2 = H 2 3 model with phantom dark energy 4 model with dynamical E.Q.S 5 quintessence model 0 nΩm,0(1 + z)3 + (1 − Ωm,0)[AS + (1 − AS)(1 + z)3(1+α)] 0 {Ωm,0(1 + z)3 + (1 − Ωm,0)(1 + z)3(1+wX )} 1+z ]o 0 nΩm,0(1 + z)3 + (1 − Ωm,0)(1 + z)3(w0+w1+1) exp[− 3w1 z 0 nΩm,0(1 + z)3 + (1 − Ωm,0)(1 + z)3(1+w0(1+z)−α)o 1 1+αo H 2 = H 2 H 2 = H 2 H 2 = H 2 -- 3 -- density contains only dust matter and radiation (Freese & Lewis 2002). We also include in the analysis the bouncing model arising in the context of loop quantum gravity (the BΛCDM model) (Singh & Vandersloot 2005; Szydlowski et al. 2005) and the model with energy transfer between the dark matter and dark energy sectors (the Λ decaying vacuum model) (Szydlowski et al. 2006). We gathered above models together with their Hubble functions (with the assumption that the Universe is spatially flat) in Table 2. The main goal of this paper is to compare all these models in the light of SNIa, CMB, BAO and H(z) data. We use the Bayesian model comparison method, which we describe in the next section. This method is commonly used in the context of cosmological models se- lection (see e.g. Liddle 2004; John & Narlikar 2002; Saini et al. 2004; Parkinson et al. 2005; Mukherjee et al. 2005; Beltran et al. 2005; Mukherjee et al. 2005a; Szydlowski & Godlowski 2006; God lowski & Szyd lowski 2005; Szyd lowski et al. 2006a; Liddle et al. 2006; Liddle 2007; Sahlen et al. 2007; Serra et al. 2007; Kunz et al. 2006; Trotta 2007; Trotta 2007a). Recently the Bayesian in- formation criteria were applied in the context of choosing an adequate model of acceleration of the Universe (Davis et al. 2007). The authors showed preference for models beyond the standard FRW cosmology (so-called exotic cosmological models) whose best fit parameters reduce them to the cosmological constant model. 2. Model comparison in Bayes theory Let us consider the set of K models: {M1,· · · , MK}. In the Bayes theory the best model from the set under consideration is this one which has the largest value of the probability in the light of the data (D), so called posterior probability P (MiD) = P (DMi)P (Mi) P (D) . (1) P (Mi) is the prior probability for the model indexed by i, which value depends on our previous knowledge about model under consideration, that is to say without information coming from data D, and P (D) is the normalization constant. If we have no foundation to favor one model over another one from the set we usually assume the same values of this quantity for all of them, i.e. P (Mi) = 1 K , i = 1,· · · , K. To obtain the form of P (D) it is required that a sum of the posterior probabilities for all models from the set is equal one K Xi=1 P (MiD) = 1 −→ P (D) = K Xi=1 P (DMi)P (Mi). -- 4 -- Therefore conclusions based on the values of posterior probabilities strongly depend on the set of models and can change when the set of models is different. P (DMi) is the marginal likelihood (also called the evidence) and has the following form P (DMi) =Z L( ¯θiD, Mi)P ( ¯θiMi)d ¯θi ≡ Ei, (2) where L( ¯θiD, Mi) is the likelihood of the model under consideration, ¯θi is the vector of the model parameters and P ( ¯θiMi) is the prior probability for the model parameters. In the case under consideration we cannot obtain the value of the evidence by analytical computation. We need a numerical method or an approximation to this quantity. Schwarz (author?) (Schwarz 1978) showed that for iid observations (D = {xi}, i = 1,· · · , N) coming from a linear exponential family distribution, defined as f (xi¯θ) = exp" S Xk=1 wk(¯θ)tk(xi) + b(¯θ)# , S = d, where w1, . . . , wS, b are functions of only ¯θ ∈ Rd, t1, . . . , tS are function of only xi, the asymptotic approximation (N → ∞) to the logarithm of the evidence is given by ln E = lnL − d 2 ln N + O(1), (3) where L is the maximum likelihood and O(1) is the term of order unity in N. In this case the likelihood function has the following form i=1f (xi¯θ) = exp"N S Xk=1 wk(¯θ)Tk(D) + b(¯θ)!# , where Tk(D) = 1 i=1 tk(xi). The integral (2) can be writing as L(¯θD, M) = ΠN N PN Z exp[N g(¯θ)]P (¯θM)d¯θ, (4) where g(¯θ) = PS k=1 wk(¯θ)Tk(D) + b(¯θ). This integral has the form of the so called Laplace integral. Assume that g(¯θ) has maximum in ¯θ0 and P (¯θ0M) 6= 0. When N → ∞ exp[N g(¯θ)] will be a sharp function picked at ¯θ0. Then the main contribution to integral (4) comes from the small neighborhood of ¯θ0. In this region P (¯θM) ≈ P (¯θ0M), we can also replace g(¯θ) function its Taylor expansion around ¯θ0: g(¯θ) = g(¯θ0)− 1 2(¯θ− ¯θ0)T C −1(¯θ− ¯θ0), where [C −1]ij = h− ∂2g(¯θ) ∂θi∂θji¯θ=¯θ0 and extend the integration region to whole Rd. One can gets the asymptotic -- 5 -- of the integral (4) E = exp[N g(¯θ0)]P (¯θ0M)(cid:0) 2π N(cid:1) 2 ln N + R, where R is the term which not depend on N. One can see that N g(¯θ0) = ln L(¯θ0D, M), where ¯θ0 is the point which maximize g(¯θ) = 1 N ln L(¯θD, M), so is equivalent to ¯θM LE (the maximum likelihood estimator of ¯θ). Finally one can obtain result (3). 2 √detC and ln E = N g(¯θ0) − d d According to this result Schwarz introduced a criterion for the model selection: the best model is that which minimizes the BIC quantity, defined as BIC = −2 lnL + d ln N. (5) This criterion can be derived in such a way that it is not required to assume any specific form for the likelihood function but it is only necessary that the likelihood function satisfies some non-restrictive regularity conditions. Moreover data do not need to be independent and identically distributed. This derivation requires to assume that a prior for model parameters is not equal to zero in the neighborhood of the point where the likelihood function under a given model reaches a maximum and that it is bound in the whole parameter space under consideration (Cavanaugh & Neath 1999). It should be pointed out that an asymptotic assumption is satisfied when a sample size used in analysis is large with respect to the number of unknown model parameters. It is useful to choose one model from our models set (here indexed by s) and compare the rest models with this one. We can define ∆BICis quantity, which is the difference of the BIC quantity for the models indexed by i and s: ∆BICis = BICi − BICs and present the posterior probability in the following form P (MiD) = exp(− 1 PK k=1 exp(− 1 2 ∆BICis)P (Mi) 2∆BICks)P (Mk) . (6) Let us assume that we have computed the probabilities in the light of data D for models from the set under consideration. Then we gathered new data D1 and want to update the probabilities which we already have. We can compute probabilities in the light of new data using information coming from previous analysis, which allow us to favor one model over another: we can use posterior probabilities for models obtained in earlier computations as a prior probabilities for models in next analysis. We apply this method in evaluation the posterior probabilities for models described in the previous section using the information coming from SNIa, CMB, BAO and H(z) data. -- 6 -- 3. Application to cosmological models comparison We start with the N = 192 sample of SNIa (Riess et al. 2007; Wood-Vasey et al. 2007; Davis et al. 2007). In this case the likelihood function has the following form L ∝ exp"− 1 2 N Xi=1 (µtheor i )2 i − µobs σ2 i !# , where σi is known, µobs of SNIa), µtheor luminosity distance, which with assumption that k = 0 is given by i = mi − M (mi -- the apparent magnitude, M -- the absolute magnitude = 5 log10 DLi + M, M = −5log10H0 + 25 and DLi = H0dLi, where dLi is the i dLi = (1 + zi)cZ zi 0 dz′ H(z′) . In this case we used the BIC quantity as an approximation to the minus twice logarithm of evidence and assumed that all models have equal values of prior probabilities. Based on our previous experiences we have used following assumptions for models parameters values: H0 ∈ h60, 80i for all models and additional: • model 2: AS ∈ h0, 1i, α ∈ h0, 1i • model 3: wX ∈ h−4,−1) • model 4: w0 ∈ h−3, 3i, w1 ∈ h−3, 3i • model 5: w0 ∈ h−3, 3i, α ∈ h0, 2i • model 7: Ωn,0 ∈ h0, 1i, n ∈ h3, 10i • model 8: Ωint,0 ∈ h−1, 1i, n ∈ h−10, 10i • model 9: n ∈ h−10, 10i • model 10: Ωl,0 ∈ h0, 4i, ΩΛb,0 ∈ h−1, 4i We separately consider cases with Ωm,0 ∈ h0, 1i and Ωm,0 ∈ h0.25, 0.31i. We treat the H0 parameter as a nuisance parameter, i.e. we marginalized the likelihood function over this parameter in the range assumed before. Posterior probabilities are obtained using equation 6. We analyse three sets of models: 1. set of models with dark energy (Table 1), 2. set of models with modified theory of gravity (Table 2), 3. set of all models (Table 1 and Table 2 together). -- 7 -- The results for the case with Ωm,0 ∈ h0, 1i are presented in Table 3, Table 4 and Table 5 for set 1, set 2 and set 3 respectively. One can conclude that in the light of SNIa data the ΛCDM is the best model from the set of models with dark energy as well as the best one from the all models under consideration, the DGP model is the best one from the group of models with modified gravity. The results for case with Ωm,0 ∈ h0.25, 0.31i are gathered in Table 6, Table 7, Table 8 for set 1, set 2 and set 3 respectively. The conclusion changed for the set of models with modified gravity: here the best one is the Cardassian model. In the next step we included information coming from CMB data. Here the likelihood function has the following form (cid:21) , H(z) dz, and Robs = 1.70± 0.03 for zdec = 1089 (Spergel et al. 2006; Wang & Mukherjee 2006). It should be pointed out that the parameter R is independent of H0. where R is so called shift parameter, Rtheor =pΩm,0R zdec L ∝ exp(cid:20)− H0 0 (Rtheor − Robs)2 2σ2 R The values of the evidence were obtained by the numerical integration. We assumed flat prior for all model parameters. It is known that evidence depends on the prior probabilities for model parameters. Assumptions for the model parameters intervals, which we made in previous analysis could be not appropriate here. Due to this we made a stricter analysis for models with parameters which interval width exceeds one. This width was used for convenience. We computed the evidence for different parameter intervals, which do not exceed the range assumed before and with a minimal width equal to one. There are of course extremely many possibilities. We limited our analysis to intervals ha, bi, where a and b are integer. Finally we chose the case with the greatest evidence. We consider the situation with Ωm,0 ∈ h0, 1i and Ωm,0 ∈ h0.25, 0.31i. The range for parameters which change after stricter analysis in the first case: • model 3: wX ∈ h−2,−1) • model 4: w0 ∈ h−1, 0i, w1 ∈ h−2, 0i • model 5: w0 ∈ h−3,−2i, α ∈ h1, 2i • model 7: Ωn,0 ∈ h0, 1i, n ∈ h3, 4i • model 8: Ωint,0 ∈ h−1, 0i, n ∈ h−10,−9i -- 8 -- • model 9: n ∈ h0, 1i • model 10: Ωl,0 ∈ h0, 1i, ΩΛb,0 ∈ h0, 1i and in the second case: • model 3: wX ∈ h−2,−1) • model 4: w0 ∈ h−1, 0i, w1 ∈ h−2,−1i • model 5: w0 ∈ h−2,−1i, α ∈ h0, 1i • model 7: Ωn,0 ∈ h0, 1i, n ∈ h3, 4i • model 8: Ωint,0 ∈ h−1, 0i, n ∈ h−2,−1i • model 9: n ∈ h0, 1i • model 10: Ωl,0 ∈ h0, 1i, ΩΛb,0 ∈ h3, 4i Posterior probabilities were obtained using equation 1. Here we treated posterior prob- abilities evaluated in analysis with SNIa data as a prior probabilities. Results are gathered in tables like in previous analysis. We also show the values of the posterior probabilities obtained for the intervals assumed at the beginning (numbers in the brackets). As we can see the ΛCDM model is still the best one from the models with dark energy (for both ranges of Ωm,0). The conclusion is the same for the set of models with modified gravity: the DGP model is the best one in the first case and the Cardassian model in the second. When we as- sume that Ωm,0 ∈ h0, 1i there is no evidence to favor the ΛCDM model over the DGP model (they have the same values of the posterior probabilities) while when we restrict Ωm,0 range to h0.25, 0.31i the ΛCDM model still stays as the best one, with even greater probability. As the third observational data we used the measurement of the baryon acoustic oscil- lations (BAO) from the SDSS luminous red galaxies (Eisenstein et al. 2005). In this case the likelihood function has the following form L ∝ exp(cid:20)− H0 (cid:17)− 1 3 h 1 H(z)i where Atheor =pΩm,0(cid:16) H(z) zA R zA H0 0 (Atheor − Aobs)2 2σ2 A (cid:21) , Here values of the evidence are obtained by the numerical integration. We made the analogous analysis with the parameters intervals as above with the additional requirement: 2 3 and Aobs = 0.469 ± 0.017 for zA = 0.35. -- 9 -- obtained intervals must at least cover the intervals obtained in previous analysis. For most of models the conclusions are the same as in the previous analysis. Below we wrote the cases where the intervals have changed for the case with Ωm,0 ∈ h0, 1i: • model 5: w0 ∈ h−3, 1i, α ∈ h1, 2i and with Ωm,0 ∈ h0.25, 0.31i: • model 8: Ωint,0 ∈ h−1, 0i, n ∈ h−2, 1i Here we used posterior probabilities obtained in analysis with the CMB data as prior probabilities. Results were again presented in described above tables. The conclusion is different for the set of all models in the case with Ωm,0 ∈ h0, 1i: the DGP model becomes the best one from them. Finally we used the observational H(z) data (N = 9) from Simon et al. (author?) (Simon et al. 2005) (see also Samushia & Ratra 2006; Wei & Zhang 2007, and references therein). These data based on the differential ages ( dt dz ) of the passively evolving galaxies a = − 1 which allow to estimate the relation H(z) ≡ a dz dt . Here the likelihood function has the following form 2" N (H(zi) − Hi(zi))2 Xi=1 L ∝ exp − #! , where H(z) is the Hubble function, Hi, zi are observational data. 1 1+z σ2 i The values of the evidence were obtained by the numerical integration. As above we assumed flat prior probabilities for models parameters. The ranges for them which changed after analogous to previous analysis are as follows: for case with Ωm,0 ∈ h0, 1i and H0 ∈ h60, 80i: • model 5: w0 ∈ h−3, 1i, α ∈ h0, 2i • model 8: Ωint,0 ∈ h−1, 0i, n ∈ h−10, 0i and for case with Ωm,0 ∈ h0.25, 0.31i and H0 ∈ h60, 80i: • model 4: w0 ∈ h−1, 0i, w1 ∈ h−2, 0i -- 10 -- Values of posterior probabilities obtained in analysis with the BAO data were treated as prior probabilities it this analysis. The results are presented in tables. As one can see in the case with Ωm,0 ∈ h0.25, 0.31i the ΛCDM model is the best one from the set of models with dark energy as well as the best one from all models considered in this paper. The conclusion is different in the set of the models with modified gravity: after the analysis with observational H(z) data the DGP model becomes the best one. In the case with Ωm,0 ∈ h0, 1i the ΛCDM is still the best model from the set of models with dark energy while the DGP model is the best one from the set of models with the modified theory of gravity as well as the best one from all models considered. As one can conclude final results coming from computation including the shrinking parameters interval procedure are the same as final results coming from computation with parameter intervals assumed at the beginning: the best model (in each set) does not change, but the probability of being the best one are greater for second case. As we have written before we used the BIC quantity as an approximation to the minus twice logarithm of the evidence for the SNIa data. This approximation gives good results if in the set under consideration is one favoured model. The problem appears when we have two favoured models with nearly the same probabilities. Such a situation is in the set of all models with Ωm,0 ∈< 0, 1 >. ΛCDM model and DGP models have nearly the same values of model probabilities which are the greatest ones in considered set. This enforce us to compute the evidence by the numerical integration for these two models (for SNIa data) to check if our previous conclusion is true. In Table 9 we gathered the values of probabilities obtained after computation of the full Bayesian evidence (case 1) as well as such values obtained when the BIC approximation was used (case 2). As one can see the conclusion changed. The ΛCDM model is better than the DGP model in the light of all data sets used in paper. The BIC approximation is not enough in this cases, gives us wrong answer. Finally we compare considered sets of models treating all described before data sets as N = 192 + 1 + 1 + 9 independent data. In this case the likelihood function has the following form L ∝ exp"− 1 2 192 Xi=1 (µtheor i )2 i − µobs σ2 i + (Rtheor − Robs)2 σ2 R + (Atheor − Aobs)2 σ2 A + Here we assumed flat prior for model parameters in the range described at the beginning of this section. We used the BIC as an approximation to −2lnE. The results for the cases with Ωm,0 ∈< 0, 1 > and Ωm,0 ∈< 0.25, 0.31 > for all described before sets of models are gathered in Tables 3, 4, 5, 6, 7, 8. (H(zi) − Hi(zi))2 σ2 i !# . 9 Xi=1 -- 11 -- Table 2. The Hubble function for cosmological models with modified theory of gravity case model H 2(z)relation DGP model BΛCDM model interacting model with Λ Cardassian model Ωr,0 = 10−4 6 7 8 9 10 Sahni-Shtanov brane I model H 2 = H 2 H 2 = H 2 0 hpΩm,0(1 + z)3 + Ωrc,0 + pΩrc,0i2ff (1−Ωm,0)2 Ωrc,0 = 4 H 2 = H 2 H 2 = H 2 0 Ωm,0(1 + z)3 − Ωn,0(1 + z)n + 1 − Ωm,0 + Ωn,0¯ 0 {Ωm,0(1 + z)3 + Ωint,0(1 + z)n + 1 − Ωm,0 − Ωint,0} H 2 = H 2 Ωr,0(1 + z)4 + Ωm,0(1 + z)42 4 0 8< : 0 nΩm,0(1 + z)3 + Ωσ,0 + 2Ωl,0 − 2pΩl,0pΩm,0(1 + z)3 + Ωσ,0 + Ωl,0 + ΩΛb,0o "0 @ 1+z + 1 1+z + (1 + z)−4+4n " 1−Ωr,0−Ωm,0 Ωm,0 1+ Ωr,0 Ωm,0 1 Ωr,0 Ωm,0 1 A n3 9= 5 ; Ωσ,0 = 1 − Ωm,0 + 2pΩl,0p1 + ΩΛb,0 Table 3. Posterior probabilities for models from Table 1; Ωm,0 ∈ h0, 1i model prior posterior SNIa posterior + CMB posterior + BAO posterior + H(z) posterior SNIa+CMB+BAO+H(z) 1 2 3 4 5 0.20 0.20 0.20 0.20 0.20 0.91 0.01 0.07 0.01 0.00 0.92 (0.95) 0.01 (0.01) 0.06 (0.04) 0.01 (0.00) 0.00 (0.00) 0.91 (0.97) 0.01 (0.01) 0.07 (0.03) 0.01 (0.00) 0.00 (0.00) 0.94 (0.99) 0.00 (0.00) 0.05 (0.01) 0.01 (0.00) 0.00 (0.00) 0.84 0.02 0.06 0.04 0.04 Table 4. Posterior probabilities for models from Table 2; Ωm,0 ∈ h0, 1i model prior posterior SNIa posterior + CMB posterior + BAO posterior + H(z) posterior SNIa+CMB+BAO+H(z) 6 7 8 9 10 0.20 0.20 0.20 0.20 0.20 0.89 0.01 0.01 0.08 0.01 0.92 (0.97) 0.00 (0.00) 0.01 (0.01) 0.07 (0.01) 0.00 (0.01) 0.92 (0.98) 0.00 (0.00) 0.01 (0.01) 0.07 (0.00) 0.00 (0.01) 0.93 (0.98) 0.00 (0.00) 0.01 (0.01) 0.06 (0.00) 0.00 (0.01) 0.07 0.03 0.13 0.74 0.03 -- 12 -- Table 5. Posterior probabilities for models from Table 1 and Table 2; Ωm,0 ∈ h0, 1i model prior posterior SNIa posterior + CMB posterior + BAO posterior + H(z) posterior SNIa+CMB+BAO+H(z) 1 2 3 4 5 6 7 8 9 10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.51 0.00 0.04 0.00 0.00 0.39 0.00 0.01 0.04 0.01 0.46 (0.48) 0.00 (0.00) 0.03 (0.02) 0.00 (0.00) 0.00 (0.00) 0.46 (0.48) 0.00 (0.00) 0.01 (0.01) 0.04 (0.01) 0.00 (0.00) 0.44 (0.46) 0.00 (0.00) 0.03 (0.01) 0.00 (0.00) 0.00 (0.00) 0.47 (0.52) 0.00 (0.00) 0.02 (0.01) 0.04 (0.00) 0.00 (0.00) 0.43 (0.45) 0.00 (0.00) 0.02 (0.00) 0.00 (0.00) 0.00 (0.00) 0.50 (0.54) 0.00 (0.00) 0.02 (0.01) 0.03 (0.00) 0.00 (0.00) 0.74 0.02 0.05 0.04 0.03 0.01 0.005 0.01 0.09 0.005 Table 6. Posterior probabilities for models from Table 1; Ωm,0 ∈ h0.25, 0.31i model prior posterior SNIa posterior + CMB posterior + BAO posterior + H(z) posterior SNIa+CMB+BAO+H(z) 1 2 3 4 5 0.20 0.20 0.20 0.20 0.20 0.91 0.01 0.06 0.01 0.01 0.98 (0.88) 0.00 (0.00) 0.01 (0.11) 0.01 (0.00) 0.00 (0.00) 0.99 (0.99) 0.00 (0.00) 0.00 (0.01) 0.01 (0.00) 0.00 (0.00) 0.99 (1.00) 0.00 (0.00) 0.00 (0.00) 0.01 (0.00) 0.00 (0.00) 0.84 0.02 0.06 0.04 0.04 Table 7. Posterior probabilities for models from Table 2; Ωm,0 ∈ h0.25, 0.31i model prior posterior SNIa posterior + CMB posterior + BAO posterior + H(z) posterior SNIa+CMB+BAO+H(z) 6 7 8 9 10 0.20 0.20 0.20 0.20 0.20 0.19 0.04 0.05 0.68 0.04 0.27 (0.26) 0.00 (0.00) 0.13 (0.04) 0.60 (0.70) 0.00 (0.00) 0.35 (0.50) 0.00 (0.00) 0.24 (0.06) 0.41 (0.44) 0.00 (0.00) 0.45 (0.91) 0.00 (0.00) 0.26 (0.04) 0.29 (0.05) 0.00 (0.00) 0.07 0.03 0.13 0.74 0.03 -- 13 -- The results in this analysis confirm that the ΛCDM model is the best one in the set of models with dark energy as well as the best one in the set of all models for both ranges in Ωm,0. The conclusion that the Cardassian model is the best one in the set of models with modified gravity (for both ranges in Ωm,0) is with contrary with previous inference where the DGP was the best one. The reason of this disagreement is related with the different constrains on the Ωm,0 parameter for various data sets for considered model. In Table 10 we presented the results of parameter estimation performed for all data sets independently as well as for all data sets applied simultaneously for the ΛCDM, DGP and Cardassian models. 4. Conclusions In this paper we gathered ten models of the accelerating Universe. The five of them explain the accelerated phase of the Universe in the term of dark energy while the other five explain this phenomenon by the modification of the theory of gravity. We used the Bayesian model comparison method to select the best one in the set of models with dark energy, in the set of models with modified theory of gravity as well as the best one of all of them. The selection based on the SNIa, CMB, BAO and observational H(z) data -- we treat posterior probabilities obtained in one analysis as prior probabilities in the next one: information coming from the previous analysis allow us to favor one model over another. We used approximation proposed by Schwarz to the minus twice logarithm of evidence in the case with SNIa data, and numerical integration of the likelihood function within an allowed parameter space (we assumed flat prior probabilities for model parameters) in the other cases. We consider separately cases with Ωm,0 ∈ h0, 1i and with Ωm,0 ∈ h0.25, 0.31i. We made a stricter analysis for models with parameters which intervals width exceed one: we evaluated the evidence for these models for different parameters intervals with minimally Table 8. Posterior probabilities for models from Table 1 and Table 2; Ωm,0 ∈ h0.25, 0.31i model prior posterior SNIa posterior + CMB posterior + BAO posterior + H(z) posterior SNIa+CMB+BAO+H(z) 1 2 3 4 5 6 7 8 9 10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.81 0.01 0.07 0.01 0.01 0.02 0.00 0.00 0.07 0.00 0.91 (0.82) 0.00 (0.00) 0.01 (0.12) 0.01 (0.00) 0.01 (0.00) 0.02 (0.02) 0.00 (0.00) 0.00 (0.00) 0.04 (0.04) 0.00 (0.00) 0.96 (0.96) 0.00 (0.00) 0.00 (0.01) 0.01 (0.00) 0.01 (0.00) 0.01 (0.02) 0.00 (0.00) 0.00 (0.00) 0.01 (0.01) 0.00 (0.00) 0.96 (0.97) 0.00 (0.00) 0.00 (0.00) 0.01 (0.00) 0.01 (0.00) 0.01 (0.02) 0.00 (0.00) 0.00 (0.00) 0.01 (0.01) 0.00 (0.00) 0.74 0.02 0.05 0.04 0.03 0.01 0.005 0.01 0.09 0.05 -- 14 -- width equal one, which do not exceed intervals assumed in the analysis with SNIa data and finally chose the best one from them (with the greatest evidence) to the next analysis. We compare such results with the results obtained in calculation where we treat all data sets as N = 192 + 1 + 1 + 9 independent data and use BIC as an approximation to the minus twice logarithm of the evidence. We can conclude that for the case with Ωm,0 ∈ h0.25, 0.31i as well as for case with Ωm,0 ∈ h0, 1i • the ΛCDM model is the best one from the set of models with dark energy as well as the best one from the set of all models considered in this paper; • the Cardassian model is the best one from the set with models with modified theory of gravity. The paper was supported by the Marie Curie Actions Transfer of Knowledge project COCOS (contract MTKD-CT-2004-517186). The authors are very grateful to the referee and to Dr W. Godlowski for helpful discussion and suggestions. REFERENCES A. G. Riess et al., AJ, 116, 1009 (1998), astro-ph/9805201. S. Perlmutter et al., ApJ, 517, 565 (1999), astro-ph/9812133. D. N. Spergel et al., ApJS, 148, 175 (2003), astro-ph/0302209. S. Weinberg, Rev. Mod. Phys. 61, 1 (1989). R. Caldwell, Phys. Lett. B545, 23 (2003), astro-ph/9908168. Table 9. Posterior probabilities for ΛCDM and DGP models model prior posterior SNIa posterior + CMB posterior + BAO posterior + H(z) ΛCDM (case1) DGP (case1) ΛCDM (case2) DGP (case2) 0.50 0.50 0.50 0.50 0.63 0.37 0.57 0.43 0.57 0.43 0.50 0.50 0.54 0.46 0.47 0.53 0.52 0.48 0.45 0.55 -- 15 -- M. Dabrowski, P. Stachowiak, M. Szydlowski, Phys. Rev. D68, 103519 (2003). M. Chevallier, D. Polarski, Int. J. Mod. Phys. D10, 213 (2001). E. V. Linder, Phys. Rev. Lett. 90, 091301 (2003). P. J. E. Peebles, B. Ratra, ApJ, L17, 325 (1988). B. Ratra, P. J. E. Peebles, Phys. Rev. D37, 3406 (1988). S. Rahvar, M. S. Movahed, Phys. Rev. D75, 023512 (2007). G. Dvali, G. Gabadadze, M. Porrati, Phys. Lett. B485, 208 (2000). Y. V. Shtanov 2000, preprint (hep-th/0005193). K. Freese, M. Lewis, Phys. Lett. B540, 1 (2002). P. Singh, K. Vandersloot, Phys. Rev. D72, 084004 (2005). M. Szydlowski, W. Godlowski, A. Krawiec, J. Golbiak, Phys. Rev. D72, 063504 (2005). M. Szydlowski, T. Stachowiak, R. Wojtak, Phys. Rev. D73, 063516 (2006). G. Schwarz, Annals of Statistics 6, 461 -- 464 (1978). J. E. Cavanaugh, A. A. Neath, Communications in Statistics -- Theory and Methods 28, 49 (1999). A. R. Liddle, MNRASL49, 351 (2004). T. M. Davis et al. 2007, preprint (astro-ph/0701510). Table 10. The best fit values of model parameters model ΛCDM SNIa CMB BAO H(z) SNIa+CMB+BAO+H(z) Ωm,0 = 0.27+0.03 −0.03 Ωm,0 = 0.22+0.04 −0.03 Ωm,0 = 0.27+0.02 −0.02 Ωm,0 = 0.31+0.1 −0.06 Ωm,0 = 0.28+0.02 −0.02 DGP Ωm,0 = 0.17+0.03 −0.02 Ωm,0 = 0.34+0.05 −0.05 Ωm,0 = 0.31+0.03 −0.03 Ωm,0 = 0.25+0.1 −0.06 Ωm,0 = 0.28+0.02 −0.02 Cardassian Ωm,0 = 0.31+0.08 −0.11 Ωm,0 = 0.26+0.32 −0.32 Ωm,0 = 0.27+0.39 −0.39 Ωm,0 = 0.01+0.16 −0.16 Ωm,0 = 0.28+0.02 −0.02 n = −0.15+0.31 −0.39 n = 0.08+1.19 −1.19 n = 0.00+3.45 −3.45 n = 0.65+0.08 −0.08 n = 0.04+0.08 −0.08 -- 16 -- M. V. John, J. V. Narlikar, Phys. Rev. D65, 043506 (2002). T. D. Saini, J. Weller, S. L. Bridle, MNRAS, 348, 603 (2004). D. Parkinson, S. Tsujikawa, B. A. Bassett, L. Amendola, Phys. Rev. D71, 063524 (2005). P. Mukherjee, D. Parkinson, P. S. Corasaniti, A. R. Liddle, M. Kunz, MNRAS, 369, 1725 (2005). M. Beltran, J. Garcia-Bellido, J. Lesgourgues, A. R. Liddle, A. Slosar, Phys. Rev. D71, 063532 (2005). P. Mukherjee, D. Parkinson, A. R. Liddle, ApJ, 638, L51 (2005a). M. Szydlowski, W. Godlowski, Phys. Lett. B633, 427 (2006). W. Godlowski, M. Szydlowski, Phys. Lett. B623, 10 (2005). M. Szydlowski, A. Kurek, A. Krawiec, Phys.Lett. B642, 171 (2006a). A. R. Liddle, P. Mukherjee, D. Parkinson, Y. Wang, Phys. Rev. D74, 123506 (2006). A. R. Liddle 2007, preprint (astro-ph/0701113). M. Kunz, R. Trotta, D. R. Parkinson, Phys. Rev. D74, 023503 (2006). R. Trotta, MNRAS, 378, 72 (2007). R. Trotta, MNRAS, 378, 819 (2007a). R. E. Kass, and A. E. Raftery, J. Amer. Stat. Assoc. 90, 773 (1995). H. Jeffreys, Theory of Probability, Oxford University Press, Oxford, third edn. (1961) A. G. Riess et al. 2007, preprint (astro-ph/9805201). W. M. Wood-Vasey et al. 2007, preprint (astro-ph/0701041). T. M. Davis et al. 2007, preprint (astro-ph/0701510). D. N. Spergel et al. 2006, preprint (astro-ph/0603449). Y. Wang, P. Mukherjee, ApJ, 650, 1 (2006), astro-ph/0604051 (2006). D. J. Eisenstein et al., ApJ, 633, 560 (2005). J. Simon, L. Verde, R. Jimenez, Phys. Rev. D71, 123001 (2005). -- 17 -- L. Samushia, B. Ratra, ApJ, 650, L5 (2006). H. Wei, S. N. Zhang, Phys.Lett. B644, 7 (2007). M. Sahlen, A. R. Liddle, D. Parkinson, Phys. Rev. D75, 023502 (2007). P. Serra, A. Heavens, A. Melchiorri 2007, MNRAS, 379, 169 (astro-ph/0701338). This preprint was prepared with the AAS LATEX macros v5.2.
0706.0074
1
0706
2007-06-01T06:38:02
Gemini Observations of Disks and Jets in Young Stellar Objects and in Active Galaxies
[ "astro-ph" ]
We present first results from the Near-infrared Integral Field Spectrograph (NIFS) located at Gemini North. For the active galaxies Cygnus A and Perseus A we observe rotationally-supported accretion disks and adduce the existence of massive central black holes and estimate their masses. In Cygnus A we also see remarkable high-excitation ionization cones dominated by photoionization from the central engine. In the T-Tauri stars HV Tau C and DG Tau we see highly-collimated bipolar outflows in the [Fe II] 1.644 micron line, surrounded by a slower molecular bipolar outflow seen in the H_2 lines, in accordance with the model advocated by Pyo et al. (2002).
astro-ph
astro-ph
Astrophysics and Space Science DOI 10.1007/sXXXXX-XXX-XXXX-X Gemini Observations of Disks and Jets in Young Stellar Objects and in Active Galaxies Peter McGregor, Michael Dopita Ralph Sutherland • Tracy Beck • Thaisa Storchi-Bergmann 7 0 0 2 n u J 1 ] h p - o r t s a [ 1 v 4 7 0 0 . 6 0 7 0 : v i X r a c(cid:13) Springer-Verlag •••• Abstract We present first results from the Near- infrared Integral Field Spectrograph (NIFS) located at Gemini North. For the active galaxies Cygnus A and Perseus A we observe rotationally-supported accretion disks and adduce the existence of massive central black holes and estimate their masses. In Cygnus A we also see remarkable high-excitation ionization cones domi- nated by photoionization from the central engine. In the T-Tauri stars HV Tau C and DG Tau we see highly- collimated bipolar outflows in the [Fe II] λ1.644 micron line, surrounded by a slower molecular bipolar outflow seen in the H2 lines, in accordance with the model ad- vocated by Pyo et al. (2002). Keywords galaxies: active -- galaxies: individual (Cygnus A, Persueus A) -- star formation: Young Stel- lar Objects -- star formation:T-Tauri stars -- Infrared: Galaxies -- Instrumentation: Adaptive Optics 1 The NIFS Instrument on Gemini North The Research School of Astronomy and Astrophysics (RSAA) of the Australian National University (ANU) recently completed the Near-infrared Integral Field Spectrograph (NIFS) which is now located at Gemini North. A preliminary design for NIFS was described by McGregor et al. (1999), and the science mission by McGregor et al. (2001). Peter McGregor, Michael Dopita Ralph Sutherland Research School of Astronomy and Astrophysics, ANU, Cotter Road, Weston Creek, ACT2611, Australia Tracy Beck Gemini Observatory, 670 N. A'ohoku Place, Hilo, HI 96720, USA Thaisa Storchi-Bergmann Instituto de Fisica, Universidade Federal do Rio Grande do Sul, Porto Alegre, RS 91501, Brasil Fig. 1. -- The data format of NIFS, the concentric im- age slicer divides the field into 29 slitlets each 0.1′′ wide with 0.04′′ pixels in the spatial direction for a total length of 3.0" on the sky. Data reduction is therefore relatively simple, since the data effectively consists of 29 long-slit spectra. 2 The NIFS instrument utilizes an innovative concen- tric image slicing design and is designed to perform near diffraction-limited, near-infrared, imaging spectroscopy with the ALTAIR facility adaptive optics system on Gemini North. The image slicer divides the field into 29 slitlets each projecting to 0.1′′ on the sky and with 0.04′′ pixels in the spatial direction. It is designed to operate in each of the J, H, and K photometric pass bands. NIFS is housed in a duplicate of the NIRI cryo- stat. It includes the Gemini On-Instrument Wavefront Sensor (OIWFS), mechanism and temperature control systems, and EPICS control software. The detector is a 2048 × 2048 HgCdTe HAWAII-2 PACE technology array manufactured by the Rockwell Science Center, developed in collaboration with the In- stitute for Astronomy of the University of Hawaii. The data format given by the NIFS instrument is illustrated in Figure 1 In summary the performance characteristics of the NIFS instrument are as follows: • 29 slitlets, each with 2040 spectral pixels, giving a total field of view of 3.0 × 3.0 arcsec. with the nucleus (be it star or AGN) used for high or- der correction and a nearby bright offset guide star used as the on-instrument wavefront sensor (OIWFS) guide star in the cases where one existd. Typically, exposures ranged between 600 and 900 sec. Equivalent length sky exposures were taken either side of these observations. Data reduction was performed in the following order using IRAF scripts especially written for NIFS. Dark subtraction, reduction including flat fielding, using a bad pixel mask to fix hot pixels, transforming the im- age by stretching in x and y based on a Ronchi mask, sky subtraction (sky images were taken in nodding pat- tern), applying a telluric correction to both object and standard stars, then using these standard stars to flux calibrate using their magnitudes as given in the 2MASS catalog. An arc was taken immediately after the object and the wavelength calibration based on this arc had a rms fitting accuracy of 0.1 Å. The final data cube ob- tained was in the form of a multi-extension FITS file with 2040 pixels in wavelength, and 29x69 spatial pix- els. • Pixel size of 0.04 x 0.10 arcsec. 3 Active Galaxies • A resolution of R ∼ 5300, corresponding to a ve- 3.1 Perseus A locity resolution of ∼ 60 km s−1 • Operation in the Z, J, H and K wavebands • Images corrected by adaptive optics (AO) • Image quality reaching down to the diffraction in H, & in J, 0.05 arcsec. limit: 0.04 arcsec. 0.06 arcsec. in K In this paper, we describe some preliminary re- sults obtained during both the commissioning time, and the guaranteed time on both pre-main sequence T-Tauri stars displaying both accretion disks and jets, and on two extragalactic sources, Cygnus A (3C 405, 4C +40.40, IRAS 19577+4035) and Perseus A (NGC 1275, 3C 084, 4C +41.07, IRAS 03164+4119), both of which show clear evidence of disks around a massive central nuclear black hole. 2 Data Acquisition & Reduction The data described below were obtained on the Gemini North telescope in commissioning time and in guaran- teed time between the dates of October 2005 and July 2006. The details will be given elsewhere. For these observations, the natural guide star system was used, The powerful radio galaxy Perseus A (NGC 1275), lo- cated at a redshift of 0.017559 (5264 km s−1), is the central component of the Perseus cluster, Abell 426. It is known by a number of other names of which the most frequently used are Perseus A, 3C 84, 4C+41.07 and IRAS 03164+4119. The cluster Abell 426 is com- posed of two distinct velocity components in the pro- cess of merging: a low velocity system moving at 5300 km s−1, which contains NGC 1275, and a high ve- locity component moving at 8200 km s−1 (Rudy et al. 1993; Krabbe et al. 2000). NGC 1275 is a star-forming early-type cD galaxy with an active nucleus. Over time, it has been assigned a whole host of classifications in- cluding peculiar Seyfert 1, Fanaroff-Riley 1 radio galaxy (F-R I), BL Lac object, a cooling flow galaxy, a LINER, and a post-merger galaxy. On the large scale, this galaxy is a strong source of thermal X-rays distributed with a complex morphology over a region of several arc minutes centred on the ac- tive nucleus (Fabian et al. 2000, 2003a). It is also sur- rounded by a large and complex filamentary Hα nebu- losity (Minkowski 1957; Lynds 1970) which has been re- cently studied in detail by Conselice, Gallagher & Wyse (2001). These filaments were believed to have been formed in a cooling flow, by the cooling of the intr- acluster medium (Fabian 1994). However, the com- plex bubble-like morphology of the X-ray gas and Gemini Observations of Disks and Jets in Young Stellar Objects and in Active Galaxies 3 the intimate connection between X-ray and Hα mor- phologies has led Fabian more recently to propose that these filaments are formed by the interaction of the radio plasma with the thermal gas in the galaxy (Fabian et al. 2003b). This picture now is much more in accord with the standard picture of the interac- tion of strong radio jets with a surrounding medium, as detailed by Bicknell, Dopita & O'Dea (1997) and Bicknell et al. (2000). The small-scale dynamical structure was first stud- ied by Wilman, Edge & Johnstone (2005), who on the basis of long-slit spectra adduced the presence of a ro- tating disk of material in the nucleus, and who on this basis, estimated a black hole mass. However, both their spatial resolution and spatial coverage were inadequate for the purposes of establishing a reliable mass. The detailed disk structure as revealed by NIFS is remarkable, and is illustrated in Figures 2 and 3. The inner portions of the molecular hydrogen and [Fe II] line images are best interpreted as a turbulent disk around the central active nucleus seen at a relatively small an- gle of inclination, and warped in its outer portions. The central inclination can be obtained from ellipse fitting of the central 0.5 arc sec of the H2 1-0 S(1) summed line emission image, which yields i = 37.1 ± 4.5◦. The surface brightness in the H2 1-0 S(1) line falls off extremely steeply with radius, and fits well to a power-law in the range 0.2 − 1.2 arc sec., or radii in the range 40 − 400pc. Such a surface brightness pro- file can be interpreted in the context of the model of Wilman, Edge & Johnstone (2005), which postulates that the molecular hydrogen emission results from tur- bulent dissipation and shocks in an unstable or clumpy accretion disk. In this, the turbulent dissipation of or- bital energy results in an increase in binding energy of the gas, resulting in a net mass inflow of the disk mate- rial. In the model, gas is repeatedly shocked as it cas- cades towards the nucleus. Therefore, the net increase of binding energy of the accretion disk as a function of radius, must be matched by radiative losses through turbulent losses and radial accretion shocks. As will be shown elsewhere (Dopita et al. 2007), the increase in the velocity dispersion of the gas towards the centre, seen in figure 3 and the observed surface bright- ness are consistent with this idea, provided that the gravitational zone of influence of the black hole is ap- proximately 0.5 arc sec. on the sky, or about 180 pc in radius. In this respect, the LINER nucleus of NGC 1275 is similar to the central disk around the central mas- sive black hole of M87, studied in detail with HST by Dopita et al. (1997). The radial velocity field can be used to constrain the enclosed mass within ∼ 35 pc, assumed here to be mass Fig. 2. -- The radial velocity field of the H2 line in Perseus A (NGC 1275). This is consistent with Kep- lerian rotation of a warped accretion disk seen almost face on around a massive central black hole. 3. -- The velocity dispersion of the H2 line in Fig. Perseus A (NGC 1275). Note how it increases sharply towards the nucleus. This is consistent with a model in which the disk is predominately heated by turbulence and shocks as the molecular material works its way to- wards the nucleus. In this model the surface brightness of the disk is predicted to increase sharply towards the nucleus, in accord with the observational results. 4 of the central black hole. However, because of the small angle of inclination, the constraints are rather poor. From the data shown in 2, Dopita et al. (2007) derive a value of log[MBH] = 8.85±0.25. This estimate is appre- ciably larger than that of Wilman, Edge & Johnstone (2005) (log[MBH] = 8.58 ± 0.18), but these authors lacked spatial coverage and were unaware of just how nearly face-on the nuclear disk in NGC 1275 actually is. 3.2 Cygnus A Cygnus A is, in bolometric terms, the most powerful radio-emitting AGN in the local universe. Indeed, its power is great enough for it to be regarded as a radio quasar. In this sense, its proximity makes it a uniquely interesting object for detailed study. On the large scale, Cygnus A displays a magnifi- cent pair of radio lobes with intense hot-spots, and very finely-collimated radio jet in the inner regions. Wilson, Smith & Young (2006) have investigated the properties of the limb-brightened, prolate spheroidal cavity in X-rays revealed by the Chandra Observatory. This cavity is filled with hotter shocked intracluster swept up by the expanding bubble of relativistic plasma associated with the radio jet. The total kinetic power of the expansion is found to be 1.2 × 1046 ergs s−1, con- sistent with the estimates of Kino & Kawakatu (2005). The circum-nuclear region shows a complex dusty bipolar structure, whose axis is very well aligned with the large-scale radio jets. This has been studied at a resolution comparable to that attained by NIFS by Canalizo et al. (2003). Their 0.05 arc sec. resolution Keck II adaptive optics near-infrared images clearly show an unresolved nucleus between two spectacular ionization/scattering cones. There is clear evidence of copious quantities of dust in these central region ac- companied with heavy nuclear obscuration. Tadhunter et al. (2000) measured the (continuum) K-band polarization and discovered > 28% polariza- tion on the nucleus and some 25% polarization in an extended linear structure aligned NNW-SSW. IR spectroscopy of the nucleus in both the H & K bands was obtained by Wilman, Fabian & Ghandi (2000), which revealed a rich emission line spectrum containing lines of molecular hydrogen, hydrogen and helium recombination lines, and forbidden lines of many species, some of which require photons of > 100eV for their excitation. For comparison with these spectra, we present NIFS spectra of the nucleus in figure 4. All of the lines identified by Wilman, Fabian & Ghandi (2000) are clearly visible. Fig. 4. -- The nuclear spectra of Cygnus A as observed in the H & K bands. Note the very high signal to noise in these spectra. Note also the existence of molecu- lar hydrogen, species of intermediate ionization and ex- tremely highly ionized species such as [S XI] along this single line-of-sight. Gemini Observations of Disks and Jets in Young Stellar Objects and in Active Galaxies 5 The spatial structure of the emission line regions ob- served with NIFS in the different ions is extremely in- structive. This is shown for some of the brighter lines in figure 5. The broad-band K continuum (Canalizo et al. 2003) shows a classical biconical structure with a cone open- ing angle of about 70 degrees. The cones are orien- tated almost exactly along the direction of the radio jet. This morphology, and the fact that the westerly cone is brighter, suggests that we are seeing the inner region of a thick torus illuminated by the central engine, seen nearly edge on, but inclined slightly such that the westerly radio jet is approaching the observer. There is also prominent point source to the SW of the nu- cleus which was suggested by Canalizo et al. (2003) to be an extragalactic source. However, this source is not apparent on any of the NIFS images, and it represents something of a mystery. The structures in the ionized species highlight differ- ent aspects of the ionization cones. The easterly cone is most prominent in the most highly excited species. This suggests that this cone is characterized by a higher ionization parameter, U, than the westerly cone. It is possible that this cone is matter bounded, giving a high [Si VI] or [Si X] to H II ratio. By contrast, in the W cone, dense matter seems to be obtruding into the cone, giving a large [Fe II] to H II ratio, but remaining rel- atively faint in the [Si X] line emission. The [Fe II] is also strong close to the nucleus, where it is presumably tracing the dense circum-nuclear gas in the inner torus. The structure in the H2 emission is remarkably dif- ferent. On the large scale it shows an extended (ring or torus?) structure extending nearly N-S. Both the align- ment and spatial extent of this structure match the the (continuum) K-band polarization structures traced by Tadhunter et al. (2000), which presumably arise from reflection of light from the central engine on the surface of dense and dusty clouds of gas. Separated from this by a low-level H2 emission, there is a bright central region elongated at right angles to the jet direction. The in- ner region presumably traces the accretion disk proper, and the outer region a disk of material, possibly in the process of being accreted to the centre. This impression is confirmed by the detailed dynamics, which are con- sistent with a slowly-rotating outer region, and a much more rapidly rotating Keplerian inner disk warped with respect to the outer ring. For Cygnus A, we may conclude that the outer struc- tures help to define the directions of escape of the ion- izing photons from the central engine, and thus the ion- ization cone structure seen in the highly ionized species. The ionization structure observed is consistent with photoionization in the radiative pressure dominated regime modelled by Groves, Dopita & Sutherland (2004a,b). It reaches very high ionization parameters in the east- erly cone; U ∼ 1.0. Finally, we can tentatively conclude from the rapid rotation seen over the central ∼ 0.2 arc sec. that the enclosed mass within ∼ 100 pc radius, assumed to be the mass of the central Black Hole, is of order log[MBH] = 9.5. 4 Young Stellar Objects (YSOs) 4.1 HV Tau C The object HV Tau C is one component of a triplet of pre-main sequence stars (Simon 1992), and displays very strong emission lines (Magazzu & Martin 1994). It represents a fine example of an accretion disk around a YSO. From high resolution near infrared images, Monin & Bouvier (2000) infer that the disk is seen nearly edge-on (i ∼ 84◦), and has a radius of ∼ 50 AU. Very similar parameters were inferred from HST imag- ing by Stapelfeldt et al. (2003), i ∼ 84◦ and R ∼ 80 AU. The disk is sufficiently inclined that (time-variable) H2O ice absorption is seen in the IR with an optical depth in the circumstellar disk of order 1.5 mag. This implies AV ∼ 20, (Terada 2005). Spectroscopically, HV Tau C has been studied in the red region of the optical spectrum, and at high resolu- tion by Apppenzeller, Bertout & Stahl (2005). They found a slight asymmetry in the [N II] lines, while the [S II] λ6716Åline is practically symmetric. In the IR, Terada (2005) made a high-resolution R = 10000 study of the [Fe II] line at 1.644µm using the Subaru IRCS and adaptive optics with a 0.3 arc sec. slit. Compo- nents at ±35km s−1 were seen, offset from one another by ∼ 0.5 arc sec., from which they inferred the existence of a jet with an opening angle of ∼ 13◦. The NIFS images stunningly reveal an [Fe II] jet em- bedded in a slower H2 conical bipolar outflow with a cone with total opening angle of about 120◦. This is shown in figure 6. The [Fe II] jet is clearly visible in the -125 km s−1 velocity channel to the E, and at +25 km s−1 to the W (velocities given are Heliocentric ra- dial velocities). For an inclination angle of only 6◦, this would imply a peak outflow velocity of ∼ 700km s−1. This is probably too high, and the true inclination an- gle is more like ∼ 10◦, which would give a true outflow velocity of ∼ 400km s−1. This picture of a strongly-collimated fast jet, prob- ably ionized by the working shocks within it, and sur- rounded by a slower, less collimated, molecular out- flow is in good accord with the model advocated by Pyo et al. (2002) for other T-Tauri outflows. 6 Fig. 5. -- The morphology of the nuclear regions of Cygnus A in various wavebands. North is at the top and east is at the left. The size of the white square is 4 × 4". The NIFS line images are integrated over ±500km s−1 equivalent waveband. Note how the molecular hydrogen image reveals the bight inner accretion disk and a system of clouds which help define the ionization cone. The [Fe II] line is confined to the circum-nuclear region and the innermost portions of the ionization cone, with a bright feature to the NW which likely represents a cloud illuminated by the central engine. Tthe remaining emission line images various aspects of the ionization cones discussed in the text. Gemini Observations of Disks and Jets in Young Stellar Objects and in Active Galaxies 7 4.2 DG Tau DG Tau B is another bright T-Tau star with strong evidence for jet outflows. It was imaged using HST and NICMOS by Padgett et al. (1999). It shows a sharply defined conical reflection region extending out to 400 AU and with a full opening angle of 80◦ on the near side, and a less clearly delineated outflow on the far side. The absorbing equatorial disk is clearly seen. The jet outflow was subjected to optical spectro- imaging observations at 0.5 arc sec. resolution with OASIS by Lavalley-Fouquet et al. (2000).This revealed a fast and unstable jet core surrounded by a slower mov- ing flow. Line ratios are consistent with shocks with speeds of 50 − 100 km s−1 with increasing flow velocity, and decreasing density away from the star. Takami et al (2004) studied the near-infrared H2 emission in DG Tau using the Infrared Camera and Spectrograph (IRCS) on the 8.2-m SUBARU telescope and found evidence for a spatial extension at right an- gles to the jet axis. The line flux ratios of the 1-0 S(0) and 1-0 S(1) lines and an upper limit for the 2-1 S(1) to 1-0 S(1) H2 line ratio suggests that the molecular flow is thermalized at a temperature of ∼ 2000 K, and is likely heated by shocks. Pyo et al. (2005) has studied the dynamics of DG Tau in the [Fe II] line at 1.644µm with an angular res- olution of up to 0.16 arc sec. achieved using the Adap- tive Optics System of Subaru Telescope. They detected two distinct velocity components separated in space and velocity. The high velocity, spatially extended compo- nents show radial velocities > 250 km s−1 while the low velocity components lie close the the exciting star and have peak velocities in the range 80 − 150 km s−1. The NIFS image in the integrated light of the [Fe II] line at 1.644µm is shown in figure 7. The jet is very clearly defined on the approaching side, but like HV Tau C, it is less well defined and has a morphology more like an expanding bubble on its receding side. On the approaching side, the velocity channels show a sharply defined core of emission, showing slight wiggles in spa- tial coordinates and displaying a smooth velocity gra- dient on the approaching side, Its velocities range from -100 km s−1 within 0.5 arc sec. of the star, and up to -300 km s−1 1.5 arc sec. away from the star. At lower velocities, and also on the approaching side, a conical feature is seen with an opening angle of about 25◦. This too displays a smooth velocity gradient ranging from - 75 km s−1 within 0.5 arc sec. of the star, up to -180 km s−1 1.5 arc sec. away. Thus there is clear dynamical evidence for two distinct [Fe II] regions - a fast jet, and a slower sheath. Fig. 6. -- The morphology of the jet and molecular outflow in HV Tau C. This image is 2 arc sec. on a side. North is at the top and east is at the left. The H2 λ2.122µm emission is shown in red. This has an outflow velocity of less than 50 km s−1, which is marginally detected in the NIFS data cube. The [Fe II] line at 1.644µm is shown in the green and the blue chan- nels. The green represents the velocity range -100 to -50 km s−1, while the blue represents the velocity range -50 to 0 km s−1. Note the bi-conical H2 region around the much faster expanding and well-collimated jet on the W side which has a full opening angle < 30◦. The east- ern jet is apparently disrupted and forms a bubble-like structure. 8 The receding lobe (where it is not obscured by the central equatorial disk) looks much more like an ex- panding bubble of gas, except for an axial enhancement in brightness seen at velocities +100 to +260 km s−1, which probably represents shocks in the receding jet. The smooth acceleration of both the jet and its sheath require that there be a source of energy in the outflowing gas, almost certainly stored magnetic energy. Thus, we agree with Lavalley-Fouquet et al. (2000) that the jet must be dominated by its magnetic energy, as proposed in the X-wind model of Shu et al. (1994), the disk models of Küker, Henning & Rüdiger (2004) or the ambipolar diffusion model of Garcia et al. (2001). These classes of model are reviewed by Gómez de Castro (2004). These magnetic jet mod- els offer a natural way to explain the (otherwise puz- zling) observations of a soft X-ray jet in DG Tau A by Güdel et al. (2004), which can be interpreted as result- ing from thermal emission from shocks in the turbulent sheath of an accelerating magnetically-dominated jet. We are undertaking hydrodynamical modelling of such jets using the new Fyris Alpha code developed by Sutherland. This is a 3D multi-level refined mesh PPM code which uses a nested and adaptive grid with Lagrangian re-map. It includes both atomic and molec- ular radiation, both time and space variable equation of state and ionisation, and self - gravity. The initial computational and physical parameters determined for the DG Tau jet system are as follows: • Computation box: 1016 cm on a side with 3 levels of resolution giving a maximal resolution of 7.72× 1012 cm, or 0.51 AU. • Gravity: Fixed isothermal spherical self-gravity potential with < V >= 4 km/s and Rcore = 2.0 × 1015 cm (or 133 AU). This potential puts a mass of M = 2.81M⊙ within 100 AU. • A three component disk & disk atmosphere con- sisting of a thin Keplerian disk T < 30K, a thick disk which is gravity and gas pressure supported with T ∼ 300 K and a warm atomic atmosphere with T ∼ 300 K, all components being in intial pressure equilibrium. • A "heavyÓ stellar jet with the following param- eters: outflow velocity v = 300 km s−1, radius R ∼ 1 × 1014 cm and mass-loss rate 10−12 − 10−7M⊙ yr−1. These models establish the timescales of the prob- lem, show that the observed morphologies can be re- produced in the models, and establish that there is an apparently smooth acceleration in the entrained sheath of the jet. However, these models cannot yet be con- sidered in any sense definitive. Jet acceleration occurs naturally in any supersonic flow in a divergent cavity, in this case formed by the dense disk material. This effect could be enhanced by the presence of a strong internal magnetic field. This provides an additional energy source to the expand- ing flow generating magneto-hydrodynamic accelera- tion. Neither of these mechanisms have been tested rigourously in a self-consistent simulation, and there is little theoretical work on late-time jet acceleration, as distinct from acceleration mechanisms associated with the launch region. Indeed even the launch mechanism is subject to considerable debate, although an MHD origin is the most likely hypothesis. It remains possible that the apparently steady accel- eration over the first ∼ 1.0 arc sec. of the jet flow is simply due to an observational selection effect. If the jet material was cool and intrinsically faint, only visible in small knots in the highest velocity slices, then the re- maining emission could all be due to the shocked and entrained material that is progressively accelerated to higher speeds along the jet outflow. If this is the case then the NIFs observation may have resolved the en- trainment and mass loading region of the jet, rather than the jet itself, or the actual launch region of the jet. The timescales associated with significant changes in the jet are exceedingly short, and gross morpholog- ical changes are expected within a year or two. It is therefore very important to make time-resolved series of observations on such T-Tauri jets. M.D. acknowledges the support of the ANU and the Australian Research Council (ARC) for his ARC Australian Federation Fellowship. The ARC also sup- ports M.D. & PMcG. under ARC Discovery Project DP0342844, and M.D. & R.S. under ARC Discovery Project DP0664434. References Apppenzeller, I., Bertout, C. & Stahl, O. 2005, A&A, 434, 1005 Bicknell, G. V., Dopita, M. A., & OÕDea, C.P. 1997, ApJ, 485,112. Bicknell, G. V., Sutherland, R. S., van Breugel, W. J. M., Dopita, M. A., Dey, A., & Miley, G. K. 2000, ApJ, 540, 678 Canalizo, G., Max, C., Whysong, D., Antonucci, R. & Dahm, S. E. 2003, ApJ, 597, 823 Conselice, C. J., Gallagher, J. S. III, & Wyse, R. F. G. 2001, AJ, 122, 2281 Dopita, M.A. et al. 1997, ApJ, 490, 202 Gemini Observations of Disks and Jets in Young Stellar Objects and in Active Galaxies 9 Dopita, M.„ McGregor, P., Beasley, A., Freeman K., Wil- son, A. Storch-Bergmann, T. & R. D. Blum 2007, (in preparation) Fabian, A. C. 1994, ARA&A, 32, 277 Fabian, A. C. et al. 2000, MNRAS, 318, L65 Fabian, A. C. et al. 2003, MNRAS, 344, L43 Fabian, A. C. 2003, MNRAS, 344, L65 Garcia, P. J. V., Ferreira, J., Cabrit, S., & Binette, L. 2001, A&A, 377, 589 Gómez de Castro, A. I. 2004, Ap&SS, 292, 561 Groves, B. A., Dopita, M. A. & Sutherland, R. S. 2004, ApJS, 153, 9 Groves, B. A., Dopita, M. A. & Sutherland, R. S. 2004, ApJS, 153, 75 Güdel, M., Skinner, S. L., Briggs, K. R., Audard, M., Arzner, K., & Telleschi, A. 2004, ApJ, 626, 53 Kino, M. & Kawakatu, N. 2005, MNRAS, 364, 659 Krabbe, A et. al. 2000, A&A, 354, 439 Küker, M., Henning, Th., & Rüdiger, G. 2004, Ap&SS, 292, 599 Lavalley-Fouquet, C., Cabrit, S. & Dougados, C. 2000, A&A, 356, 41 Magazzu, A. & Martin, E. L. 1994, A&A, 287, 571 McGregor, P. J., Conroy, P., Bloxham, G., & van Harmelen, J., 1999, PASA, 16, 273 McGregor, P. J., Dopita, M., Wood, P. & Burton, M. G. 2001, PASA, 18, 41 - 57 Minkowski, R. 1957, Proc. IAU Symp 4, ed. H. C. Van de Hulst, Cambridge U.Press: Cambridge, p107 Monin, J.-L. & Bouvier, J. 2000, A&A, 356, 75 Lynds, R. 1970, ApJ, 159, L151 Padgett, D. L. et al. 1999, AJ, 117, 1490 Pyo, T.-S. et al. 2002, ApJ, 570, 724 Pyo, T.-S. et al. 2005 in Proceedings of the ESO Workshop Science with Adaptive Optics, eds W. Brandner and M. E. Kasper, Springer, Berlin 2005, p.242 Rudy, R. J. et al. 1993, ApJ, 414, 527 Shu, F., Najita, J., Ostriker, E., Wilkin, F., Ruden, S, & Lizano, S 1994, ApJ, 429, 781 Simon, M., Chen, W. P., Howell, R. R., Benson, J. A., & Slowik, D. 1992, ApJ, 384, 212 Stapelfeldt, K. R et al. 2003, ApJ, 589, 410 Tadhunter, C. N. et al. 2000, MNRAS, 313, 52 Takami, M et al. 2004, A&A, 416, 213 Terada, H. 2005, poster presented at the Conference "Pro- tostars & planets V" Wilman, R. J., Fabian, A. C., & Ghandi, P. 2000, MNRAS, 318, 1232 Wilman, R. J., Edge,A. C. & Johnstone, R. M. 2005, MN- RAS, 359, 755 Wilson, A. S., Smith, D. A. & Young, A. J. 2005, ApJ, 644, 9 This 2-column preprint was prepared with the AAS LATEX macros v5.2. Fig. 7. -- The morphology of the DG Tau jet as seen by NIFS in the integrated light of the [Fe II] line at 1.644µm. North is at the top east at the left, and the image is 3 × 3arc sec in size, which corresponds to 460 AU. Note that the (inclined) equatorial disk is clearly visible, and is therefore still optically thick at 1.644µm within a projected distance of about 100 AU .
astro-ph/0509912
1
0509
2005-09-30T16:20:58
Supermassive Black Hole Mass Functions at Intermediate Redshifts from Spheroid and AGN Luminosity Functions
[ "astro-ph" ]
Redshift evolution of supermassive black hole mass functions (BHMFs) is investigated up to z ~ 1. BHMFs at intermediate redshifts are calculated in two ways. One way is from early-type galaxy luminosity functions (LFs); we assume an M_BH - L_sph correlation at a redshift by considering a passive evolution of L_sph in the local relationship. The resultant BHMFs (spheroid-BHMFs) from LFs of red sequence galaxies indicates a slight decrease of number density with increasing redshift at M_BH > 10^{7.5-8} M_solar. Since a redshift evolution in slope and zeropoint of the M_BH - L_sph relation is unlikely to be capable of making such an evolution in BHMF, the evolution of the spheroid-BHMFs is perhaps due mainly to the decreasing normalization in the galaxy LFs. We also investigate how spheroid-BHMFs are affected by uncertainties existing in the derivation in detail. The other way of deriving a BHMF is based on the continuity equation for number density of SMBHs and LFs of active galactic nucleus (AGN). The resultant BHMFs (AGN-BHMFs) show no clear evolution out to z = 1 at M_BH > 10^8 M_solar, but exhibit a significant decrease with redshift in the lower mass range. Comparison of the spheroid-BHMFs with the AGN-BHMFs suggests that at M_BH > 10^{8} M_solar, the spheroid-BHMFs are broadly consistent with the AGN-BHMFs out to z ~ 1. The agreement between the spheroid-BHMFs and the AGN-BHMFs appears to support that most of the SMBHs are already hosted by massive spheroids at z ~ 1 and they evolve without significant mass growth since then.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 13 (2005) Printed 13 June 2021 (MN LATEX style file v2.2) Supermassive Black Hole Mass Functions at Intermediate Redshifts from Spheroid and AGN Luminosity Functions Naoyuki Tamura1⋆, Kouji Ohta2, & Yoshihiro Ueda2,3 1Department of Physics, University of Durham, South Road, Durham, DH1 3LE, UK 2Department of Astronomy, Kyoto University, Kyoto 606-8502, Japan 3Institute of Space and Astronautical Sciences, 3-1-1 Yoshinodai, Sagamihara-shi, Kanagawa 229-8510, Japan ABSTRACT Redshift evolution of supermassive black hole mass functions (BHMFs) is in- vestigated up to z ∼ 1. BHMFs at intermediate redshifts are calculated in two ways. One way is from early-type galaxy luminosity functions (LFs); we assume an MBH − Lsph correlation at a redshift by considering a passive evolution of Lsph in the local relationship. The resultant BHMFs (spheroid-BHMFs) from LFs of red se- quence galaxies indicates a slight decrease of number density with increasing redshift at MBH > 107.5−8M⊙. Since a redshift evolution in slope and zeropoint of the MBH −Lsph relation is unlikely to be capable of making such an evolution in BHMF, the evolution of the spheroid-BHMFs is perhaps due mainly to the decreasing normalization in the galaxy LFs. We also derive BHMFs from LFs of morphologically selected early-type galaxies. The resultant BHMFs are similar to those from the red sequence galaxies, but show a small discrepancy at z ∼ 1 corresponding to an increase of SMBH num- ber density by ∼ 0.3 dex. We also investigate how spheroid-BHMFs are affected by uncertainties existing in the derivation in detail. The other way of deriving a BHMF is based on the continuity equation for number density of SMBHs and LFs of active galactic nucleus (AGN). The resultant BHMFs (AGN-BHMFs) show no clear evolution out to z = 1 at MBH > 108M⊙, but exhibit a significant decrease with redshift in the lower mass range. Interestingly, these AGN- BHMFs are quite different in the range of MBH 6 108M⊙ from those derived by Merloni (2004), where the fundamental plane of black hole activity is exploited. Comparison of the spheroid-BHMFs with the AGN-BHMFs suggests that at MBH > 108M⊙, the spheroid-BHMFs are broadly consistent with the AGN-BHMFs out to z ∼ 1. Although the decrease of SMBH number density with redshift suggested by the spheroid-BHMFs is slightly faster than that suggested by the AGN-BHMFs, we presume this to be due at least partly to a selection effect on the LFs of red sequence galaxies; the colour selection could miss spheroids with blue colours. The agreement between the spheroid-BHMFs and the AGN-BHMFs appears to support that most of the SMBHs are already hosted by massive spheroids at z ∼ 1 and they evolve without significant mass growth since then. Key words: black hole physics - galaxies: elliptical and lenticular, cD -- galaxies: evolution. 1 INTRODUCTION Recent observations provide evidence that a mass of a super- massive black hole (SMBH) in a galactic nucleus is tightly correlated with a mass or luminosity of a spheroid compo- nent of its host galaxy (e.g., Magorrian et al. 1998; Mar- coni & Hunt 2003, MH03 hereafter). The tight correla- ⋆ E-mail:[email protected] tion suggests the presence of strong evolutionary link be- tween SMBH and spheroid component. Using the relation, an SMBH mass function (BHMF) can be derived from lo- cal galaxy luminosity function (LF) or velocity dispersion function. Meanwhile, a local BHMF can also be calculated from cosmological evolution of LFs of active galactic nuclei (AGNs) by integrating the continuity equation for number density of SMBHs, where mass accretion onto a SMBH is as- sumed to power an AGN and grow the central SMBH (e.g., 2 N. Tamura, K. Ohta, & Y. Ueda Cavaliere, Morrison & Wood 1971; Small & Blandford 1992; Marconi et al. 2004; Shankar et al. 2004). BHMFs derived by this method can now be more reliable than before thanks to updated LFs of hard X-ray selected AGNs (Ueda et al. 2003), which is more complete to obscured AGNs. Marconi et al. (2004) and Shankar et al. (2004) demonstrate that the local densities of SMBHs and BHMFs derived with these two ways agree with each other, if one adopts reasonable values for accretion efficiency and Eddington ratio. Furthermore, Marconi et al. (2004) indicate that the cosmic history of mass accretion rate density delineates that of star-formation rate density, and the ratio of the latter to the former is about 4000, which agrees with the mass ratio of a spheroid to a central SMBH, again strongly suggesting the co-evolution of SMBHs and spheroids. In studying the co-evolution of SMBHs and host spheroids in further detail, one approach is to investigate redshift evolution of BHMF and correlation between black hole mass (MBH) and spheroid luminosity (Lsph) or mass. While MBH has been measured for a substantial number of high redshift QSOs (e.g., Shields et al. 2003; McLure & Dunlop 2004), it is technically challenging to directly mea- sure dormant SMBHs at cosmological distances. However, BHMFs at high redshifts can be computed using AGN LFs and the continuity equation in the same way as in the lo- cal universe. Also, we are now in a reasonable position to be able to study BHMFs at intermediate redshifts by using LFs of early-type galaxies. Early-type galaxy LFs have recently been derived out to z ∼ 1 with good statistics from inten- sive imaging surveys such as COMBO-17 (Wolf et al. 2003; Bell et al. 2004b). Consequently, one can compare BHMFs from the galaxy LFs with those from AGN LFs, which may provide clues to understand the co-evolution of AGNs and host spheroids. In this paper, we will derive BHMFs at intermediate redshifts from galaxy LFs and AGN LFs and investigate the redshift evolutions. When converting galaxy LFs to BHMFs, a correlation between MBH and Lsph is utilized. Although a correlation between MBH and bulge effective stellar ve- locity dispersion (σe) is claimed to be tighter than that be- tween MBH and Lsph (Ferrarese & Merritt 2000; Gebhardt et al. 2000), measuring σe at cosmological distance is very hard and we thus adopt the MBH − Lsph relation in this work. It should also be emphasized that MH03 show the MBH − Lsph relation is as tight as that between MBH and σe if only SMBHs whose masses are securely determined are used in the analysis. In addition, MH03 suggest that the in- trinsic scatter of the relation in B-band is as small as in NIR bands. One should be aware that the derivation of a BHMF from galaxy LFs contains several processes which may give significant uncertainties to resultant BHMFs. Quantifying uncertainties in BHMFs when calculated from galaxy LFs is also aimed at in this paper. The layout of this paper is as follows. In the next section (§ 2), we describe the procedure to derive a BHMF from an early-type galaxy LF and calculate BHMFs from early-type galaxy LFs up to z ∼ 1. We also investigate how BHMFs are affected by uncertainties and unconstrained parameters existing in this derivation. In § 3, we derive BHMFs from AGN LFs and examine its redshift evolution. In § 4, we compare the BHMFs from early-type galaxy LFs with those from AGN LFs and discuss the results. Throughout this pa- per, we adopt the cosmological model with H0 = 70 km s−1 Mpc−1, ΩM = 0.3 and ΩΛ = 0.7 unless otherwise stated. 2 BHMF FROM EARLY-TYPE GALAXY LF 2.1 Derivation of BHMF In this paper, we mainly use the LFs obtained by Bell et al. (2004b) from the COMBO-17 survey. The large survey area with the moderate depth brings a large number of galaxies, and the multi-band photometry covering from 3640 A to 9140 A with the 17 broad- and medium-band filters allows one to accurately determine the photometric redshifts. Both of these aspects are important to derive luminosity functions with good accuracy and the large survey area is especially useful to estimate its cosmic variance. They firstly investi- gated the rest-frame U − V vs. MV colour-magnitude di- agram of galaxies at a certain redshift and found the red sequence consistent with the colour-magnitude relation well established for the early-type galaxy population. They select galaxies on the red sequence and derive their rest-frame B- band LFs at redshifts from 0.25 to 1.05. Bell et al. (2004b) also derived the B-band LF of the local red sequence galax- ies using the SDSS EDR data (Stoughton et al. 2002) by transforming the SDSS ugr system to the standard U BV system. We use the Schechter functions fitted to these LFs in the following analyses. A part of the survey area of the COMBO-17 was imaged with the HST/ACS and most of the galaxies on the red sequence (∼ 85 %) indeed show early- type morphology (Bell et al. 2004a). It should be noted that the colour selection based on the red sequence is presumed to exclude blue ellipticals with star formation activity and/or young stellar population. In addition, small bulges in late- type galaxies can be missed, which perhaps results in a de- ficiency of the light part of a BHMF. In order to convert these LFs to BHMFs, firstly the total galaxy luminosities need to be transformed to the spheroid luminosities using bulge-to-total luminosity ratios (B/T s). B/T s of early-type galaxies at intermediate red- shifts are, however, not well constrained observationally. Im et al. (2002) selected morphologically early-type galax- ies at intermediate redshifts (most of the galaxies are at z = 0.2−1.0) from the HST/WFPC2 data for DEEP Groth- Strip Survey (DGSS) based on B/T s (> 0.4), which are estimated by fitting radial surface brightness profiles with r1/4 spheroid and exponential disk. According to their cata- log, the morphologically selected early-type galaxies have a B/T of 0.7 on average, although there is a substantial scat- ter. The B/T s do not depend on redshift or luminosity in their sample. Therefore we adopt B/T = 0.7 in this study, independently of redshift and galaxy luminosity. Once LFs of spheroidal components are calculated from the early-type galaxy LFs, they can be transformed to BHMFs using an MBH − Lsph relation. In this study, we use those derived by MH03 for their galaxies in Group 1, for which measurements of MBH and Lsph are considered to be reliable. The relation is derived for Lsph in the B, J, H, and K bands. Although the relation in K-band is generally preferred because K-band luminosity is considered to be the most reliable indicator of the stellar mass of a galaxy, MH03 show the intrinsic scatter of the relation is almost indepen- dent of the bands for the galaxies in Group 1. Since we will Supermassive black hole mass functions at intermediate redshifts 3 mainly use the B-band LFs by Bell et al. (2004b), we adopt the B-band relation described as follows: log MBH = (1.19 ± 0.12)(log Lsph − 10.0) + (8.18 ± 0.08).(1) In applying the MBH − Lsph relation to an LF, the intrin- sic scatter of the relation needs to be considered; we adopt ∆ log MBH = 0.32 according to MH03. In calculating a BHMF from a spheroid LF at a high redshift, an evolution of the MBH − Lsph relation needs to be considered. In this study, we basically consider only the effect of passive evolution in Lsph. It should be noted that if all spheroid components evolve only passively (with no growth of SMBHs), then the BHMF at the intermediate redshift does not change from that at z = 0. In fact, it is not obvious that all spheroids evolve passively at intermediate redshifts. Any other evolutions in galaxy LFs and the MBH − Lsph relation would cause the disagreement between BHMFs from galaxy LF and AGN LF, which will be discussed in § 4. We describe the passive luminosity evolution as MB(z) = MB(z = 0) − Qz and assume Q to be 1.4. This is evaluated using PEGASE Ver 2.0 (Fioc & Rocca-Volmerange 1997) for a stellar population with the solar metallicity formed at z = 4 (the age at z = 0 is 12 Gyr) in a single starburst with an e-folding time of 1 Gyr. This Q value is consistent with the evolution of characteristic luminosity in the COMBO-17 LFs (z > 0.25). 2.2 BHMFs from early-type galaxy LFs up to z ∼ 1 Figure 1 shows the BHMFs calculated using the COMBO- 17 LFs with the prescription described above. In the upper panel, the B-band Lsph of the galaxies in Group 1 by MH03 is plotted against MBH. Black solid line indicates the best- fitting regression line to the data (i.e., the relation at z = 0). Grey solid, black dashed, and black dot-dashed line is the relation expected at z = 0.25, 0.65 and 1.05, respectively, when the passive luminosity evolution is considered. In the lower panel, BHMFs transformed from the COMBO-17 LFs in the rest-frame B band at these redshifts are indicated with shaded regions. The envelope of each BHMF shows the errors of M ∗ B and φ∗ in the Schechter function fit to the data (Bell et al. 2004b); the upper and lower bound is defined by a BHMF with the largest and smallest L∗ B and φ∗ within the fitting error, respectively. The uncertainty in φ∗ is dominated by cosmic variance (Bell et al. 2004b) and we adopt the larger value of the two different estimates by Bell et al. (2004b). Each BHMF is indicated down to the black hole mass corresponding to the lowest luminosity among the data points in the LF. The characteristic MBH corresponding to M ∗ B at each redshift is indicated by arrow with the same colour as the BHMF. The BHMFs in Figure 1 exhibit a redshift evolution. Since the characteristic MBH does not largely change with redshift, this evolution is perhaps due to a density evolution of BHMFs corresponding to the decreasing normalization of the red sequence galaxy LFs (Bell et al. 2004b). We will discuss other possibilities in § 4.2. Currently, the COMBO-17 survey provides one of the largest and most useful databases to study galaxies out to z ∼ 1 and the derived early-type galaxy LFs are there- fore considered to be the most reliable so far. Neverthe- Figure 1. Upper panel: The correlation between MBH and B- band spheroid luminosity. Black solid line indicates the relation at z = 0 (equation (1)) fitted to the data points by MH03 for their galaxies in Group 1. Grey solid, black dashed, and dotted line is the relation expected at z = 0.25, 0.65 and 1.05, respectively, when a passive evolution of spheroid luminosity is considered. Lower panel: BHMFs transformed from the COMBO-17 LFs at redshifts of 0.25, 0.65, and 1.05 are indicated with shaded regions, B and φ∗ (Bell et al. 2004b). of which widths show the errors in M ∗ The lower mass cutoff of these BHMFs corresponds to the lowest luminosity of the data points in the original LF. The characteristic MBH corresponding to M ∗ B at each redshift is indicated by arrow with the same colour as of the BHMF. less it is worth investigating the evolution of BHMF using LFs determined with other data sets, in particular to see whether the result is sensitive to selection criterion for early- type galaxy. Since early-type galaxies are selected from the colour-magnitude diagram in COMBO-17, we investigate the BHMFs derived from LFs of morphologically selected early-type galaxies by Im et al. (2002; the HST/WFPC2 data for DGSS are used) and by Cross et al. (2004; the data were taken with the HST/ACS in the guaranteed time observations). In these studies, early-type galaxies are se- lected based on the analyses of radial surface brightness profiles. In Im et al. (2002), the rest-frame B-band LFs are derived in the two redshift bins: 0.05 < z < 0.6 and 0.6 < z < 1.2, while in Cross et al. (2004), they are ob- tained at 0.5 < z < 0.75 and 0.75 < z < 1.0. In Figures 2 and 3, the BHMFs converted from the LFs of morphologically selected early-type galaxies by Im et al. (2002) and Cross et al. (2004), respectively, are presented by shaded regions, showing the uncertainties due to the fit- ting errors of M ∗ B and φ∗ in the LFs. Again these BHMFs are calculated by assuming B/T = 0.7 and using the B- band MBH − Lsph relation by MH03 corrected for passive luminosity evolutions. The low mass cutoff of the BHMF corresponds to the lowest luminosity among the data points in the original LF. The BHMF from the COMBO-17 LF at a 4 N. Tamura, K. Ohta, & Y. Ueda Figure 2. The BHMFs converted from the early-type galaxy LFs in the DGSS (Im et al. 2002) are indicated with shaded regions, whose widths represent the uncertainties of the BHMFs due to B and φ∗ in the LFs. The pair of solid lines the fitting errors of M ∗ describes the BHMF from COMBO-17 at similar redshifts and the separation of the two lines indicates the uncertainty of the BHMF. The low mass cutoff of the BHMF corresponds to the lowest luminosity among the data points in the LF. Figure 3. Same as Figure 2, but for the BHMFs converted from the early-type galaxy LFs computed by Cross et al. (2004) using the HST/ACS data. similar redshift is also indicated with solid lines, showing the upper and lower bounds defined by considering the errors of M ∗ B and φ∗ in the LF. These comparisons demonstrate that the BHMFs from the morphologically selected early-type galaxy LFs are consistent with those from the COMBO-17 survey. It should be mentioned that the BHMFs obtained from COMBO-17 LFs tend to lie below those from the mor- phologically selected early-type galaxy LFs at the higher redshifts; the discrepancy in number density of SMBHs at a given MBH is estimated to be ∼ 0.3 dex. This may indicate that the contribution of blue spheroids which are not in- cluded in the COMBO-17 LFs becomes larger towards z ∼ 1. 2.3 Source of uncertainty in the derivation of a BHMF In the previous subsections, we showed the procedure to convert galaxy LFs to BHMFs at intermediate redshifts and presented the resultant BHMFs up to z ∼ 1. However, there are several sources of uncertainty in the derivation. Some of them are related to the MBH − Lsph relation such as fitting error of the relation to the data points, uncertainty of its intrinsic scatter, and choice of MBH − Lsph relations. Others are due to the fact that B/T value and passive luminosity evolution are not well constrained. In what follows, we will investigate how BHMFs are affected by these uncertainties. Unless otherwise noted, we will begin all the calculations to derive BHMFs with the LF of red sequence galaxies at z = 0 (Bell et al. 2004b, the SDSS EDR data are used). We call this input LF "iLF" hereafter. 2.3.1 Fitting error and uncertainty of intrinsic scatter in MBH − Lsph relation In Figure 4, the fitting error in the B-band MBH − Lsph re- lation estimated by MH03 is demonstrated in the top panel. Three relations are shown here: One is the best-fitting re- gression line (solid line), and the other two are those which result in the most massive or least massive BHMFs within the ±1σ fitting error. The BHMF using either of the three relations is indicated in the middle panel. In calculating the BHMFs, the intrinsic scatter ∆ log MBH = 0.32 (MH03) is taken into account. This shows that the part of a BHMF at MBH > 108.5M⊙ is affected by this uncertainty. We also calculate BHMFs for several values of intrinsic scatter of the relation and show the results in the bottom panel; 0 (grey solid line), 0.2 (dotted line), 0.3 (black solid line), and 0.4 (dashed line). This affects again the massive end of a BHMF (MBH > 108.5M⊙). It is suggested by MH03 to be between 0.3 and 0.4 (see also McLure & Dunlop 2002), but we note that if observational errors in MBH are under- estimated, the intrinsic scatter would be smaller. 2.3.2 MBH − Lsph relations in B band and K band We will compare BHMFs derived from iLF using the MBH − Lsph relations calculated by MH03 either in the B band or K band. In order to apply the K-band relation to iLF (B band), we need to convert the K-band relation to a B- band relation by correcting it for B − K colour. This can be performed by using the average colour of early-type galaxy Supermassive black hole mass functions at intermediate redshifts 5 Figure 5. Upper panel: B-band MBH −Lsph relations. Black solid line is the B-band MBH − Lsph relation by MH03 and the other lines indicate those obtained by correcting the K-band relation for A-CMR (grey solid line), T-CMR (dotted line), and the average B − K colour (dashed line). Lower panel: BHMFs calculated from iLF using the above B-band relations are indicated. However, this correction ignores the colour-magnitude relation (CMR). In fact, Bell et al. (2004b) suggest that the red sequence galaxies in the COMBO-17 survey are on the CMR which is consistent with those found for E/S0s both in clusters (Bower, Lucey, & Ellis 1992; Terlevich, Caldwell & Bower 2001) and fields (Schweizer & Seitzer 1992). There- fore, we attempt to take this into account. Since CMRs in B −K have rarely been investigated, we model a CMR using a population synthesis code (Kodama & Arimoto 1997) so as to reproduce an observed CMR in a certain set of filters and we then derive a CMR in B − K and MB using the model. In this modelling calculation, we follow the recipe by Kodama et al. (1998), where they modeled the CMR (V −K and MV ) measured in the Coma cluster (Bower et al. 1992) based on the galactic wind scenario. It needs to be mentioned that, while a CMR is normally defined using colours within a metric aperture, the one us- ing total colours suits better for this study. Total colours in more luminous (larger) galaxies are presumed to be system- atically bluer than those measured within an aperture due to more severe effects of the colour gradients (e.g., Peletier et al. 1990; Tamura & Ohta 2003); a given aperture can sample only the reddest part of a luminous (large) galaxy, but it can include the total light of a faint (small) galaxy. The slope of a CMR using total colours (T-CMR hereafter) is therefore expected to be flatter than that using colours within an aperture (A-CMR hereafter). This aperture effect on the CMR in the Coma cluster (Bower et al. 1992) has been investigated by Kodama et al. (1998) and the slope of the T-CMR at MV 6 −20 mag is estimated to be ∼ 30% Figure 4. Top panel: The MBHs and Lsphs of the galaxies in Group 1 by MH03 are plotted and the best-fitting regression line to the data points is indicated by solid line. Dashed (dotted) line shows the relation which results in a BHMF most (least) biased towards the massive end within the 1 σ fitting errors, respectively, when applied to an LF. Middle panel: BHMFs are computed from iLF with the MBH − Lsph relations demonstrated above and they are indicated with the same line styles as in the top panel. The intrinsic scatter of the relation (∆ log MBH = 0.32; MH03) is considered to derive the BHMFs. Bottom panel: BHMFs are cal- culated for several values of intrinsic scatter of the relation; 0 (grey solid line), 0.2 (dotted line), 0.3 (black solid line), and 0.4 (dashed line). population (e.g., Marconi et al. 2004; McLure & Dunlop 2004; Shankar et al. 2004). We adopt B−K = 3.75 estimated by Girardi et al. (2003) for local elliptical and S0 galaxies in field and group environments. This colour is similar to that used in Marconi et al. (2004), where MZ − K = 4.1 for ellipticals and 3.95 for S0s (MZ is Zwicky magnitude) are used in applying the K-band MBH − Lsph relation to the LFs from the CfA survey (Marzke et al. 1994). These colours are based on actual measurements by Kochanek et al. (2001) and they are converted to B − K = 3.59 and 3.44, respectively, by using the equation B = MZ − 0.51 given by Aller & Richstone (2002). 6 N. Tamura, K. Ohta, & Y. Ueda flatter. Hence we consider a CMR with a slope flatter by 30% than the A-CMR by Bower et al. (1992) as a T-CMR.1 In Figure 5, the B-band MBH −Lsph relations are shown in the upper panel. Black solid line is the B-band MBH−Lsph relation by MH03 and the other lines indicate those obtained by correcting the K-band relation for the average B − K colour, A-CMR, and T-CMR. In the lower panel, BHMFs calculated from iLF using these B-band relations are in- dicated with the same line styles as above. It is indicated that the K-band relation is not fully transformed to the B-band relation by MH03 even if a CMR is taken into ac- count, and the BHMFs come towards the less massive end than the case where the B-band relation by MH03 is ap- plied to iLF. This may imply that, while the tightness of the B-band MBH − Lsph relation is nearly the same as that of the K-band relation, the relations are not equivalent to each other. One possible reason for this discrepancy may be that B band luminosity is less good indicator of stellar mass due to effects of dust extinction and/or young stellar pop- ulation for some of the galaxies used in the analysis of the MBH − Lsph relations. It should be noted that the BHMF comes towards the less massive end when the average B − K colour is used than when a CMR is considered because the adopted average colour is bluer than colours of luminous early-type galaxies. Consequently, their B-band luminosi- ties are overestimated and therefore smaller values of MBH are assigned to spheroids with a certain B-band luminosity. For the same reason, the BHMF also depends on choice of A-CMR or T-CMR, but the difference turns out to be very small. 2.3.3 Choice of MBH − Lsph relations from different authors Next we will compare BHMFs derived using the MBH −Lsph relations by different authors: Ferrarese & Merritt (2000, FM00), Merritt & Ferrarese (2001, MF01), McLure & Dun- lop (2002, MD02)2, and MH03. For FM00, we adopt the re- lation for their Sample A. In MF01 and MD02, the relation is obtained in the V band and R band, respectively, and they are converted to B-band relations using T-CMRs modeled in the same way as explained earlier. Also, the zeropoints of the relations by MF01 and MD02 are shifted by the amounts 1 In Kodama et al. (1998), H0 = 50 km s−1 Mpc−1 is assumed and hence the slope of a T-CMR would be less flat in our cos- mology (H0 = 70 km s−1 Mpc−1). But here we aim at seeing how BHMFs vary by considering the CMRs and precise deter- mination of the slope is beyond our scope. We note that colours of ellipticals less luminous than MV = −20 mag are considered to be robust to the aperture correction and the actual T-CMR therefore has a break at MV ∼ −20 mag (see Figure 3 in Kodama et al. 1998). We ignore this and determine the zeropoint of the T-CMR so as to reproduce the total colours of the ellipticals with MV 6 −20 mag. Although this indicates that the T-CMR gives too red colours to less luminous ellipticals, the impact of this on a BHMF is very small and does not affect the following discussions. 2 Strictly speaking, since about half of the sample consists of QSOs at 0.1 < z < 0.5 in MD02, the MBH − Lsph relation is not allowed to be used here because some evolutionary effects may already be incorporated. In practice, however, the MBH − Lsph relation derived using only the local inactive galaxies in MD02 is the same as that from the whole sample. Figure 6. Same as Figure 5, but MBH−Lsph relations by different authors and BHMFs derived using them are compared. Upper panel: Black solid line, grey solid line, dotted line, and dashed line indicates the B-band MBH − Lsph relation by MH03, FM00, MF01, and MD02, respectively. In MD02 and MF01, the relation is originally derived in the V and R band, respectively, and they are converted to the B-band relations using T-CMR (see text for details). Lower panel: BHMFs derived by applying the above B- band relations to iLF are indicated. The intrinsic scatter of the MBH − Lsph relation is assumed to be 0.32 (MH03) in all the calculations of the BHMFs. due to the differences in H0 from the value we adopt. These B-band relations are shown in the upper panel of Figure 6 with the B-band relation by MH03. They are applied to iLF and the BHMFs obtained are shown in the lower panel. The intrinsic scatter of 0.32 around the relation is assumed in all the calculations. These BHMFs suggest that MBH − Lsph relation depends on sample data set and consequently there is a substantial variation among the BHMFs. 2.3.4 Bulge-to-total luminosity ratio B/T at intermediate redshift is only loosely constrained from observations at the moment and we have to await for future works to see the validity of the current assumption (B/T = 0.7 for all the red sequence galaxies). It is therefore worth demonstrating BHMFs for a range of B/T to keep it in mind as uncertainty. In Figure 7, BHMFs for B/T = 0.5, 0.7, and 0.9 are compared. These BHMFs are obtained from iLF and the B-band MBH − Lbulge relation by MH03. This indicates that a BHMF is affected by choice of B/T s at MBH > 108M⊙. 2.3.5 Model of passive luminosity evolution We consider a passive luminosity evolution of an old stellar population which formed at z = 4 with a starburst of which Supermassive black hole mass functions at intermediate redshifts 7 Figure 7. BHMFs for three B/T s are shown. These BHMFs are obtained from iLF and the B-band MBH −Lsph relation by MH03. Figure 9. Same as Figure 8, but for z = 1.05. at z = 0.55 and apply these relations to the COMBO-17 LF at this redshift (not iLF). The results are displayed in Figure 8. We also calculate BHMFs at z = 1.05 using the LF and the passively evolved MBH − Lsph relation at this redshift and show the results in Figure 9. 2.3.6 Summary of the uncertainties Table 1 shows a summary of uncertainties in a BHMF in- vestigated earlier. In each column, the possible range of log- arithmic SMBH number density (∆ log φM ) caused by each source of uncertainty is indicated at the five BH masses: log MBH = 7.5, 8.0, 8.5, 9.0 and 9.5. The error sources in- vestigated are arranged in the columns (1) − (6) as follows: (1) the 1σ fitting error of the B-band MBH − Lsph rela- tion (see the middle panel of Figure 4), (2) the uncertainty of intrinsic scatter (0.3 − 0.4) in the MBH − Lsph relation (the lower panel of Figure 4), (3) choice of filter bands (B or K) of MBH − Lsph relation (Figure 5), (4) choice of the MBH − Lsph relations by different authors (Figure 6) (5) B/T (0.5 − 0.9; Figure 7), and (6) choice of passive evolu- tion models (zf = 3 − 10; ∆ log φM is calculated for BHMFs at z = 1.05 (Figure 9)). In calculating ∆ log φM for (3), we compare a BHMF computed with the B-band MBH − Lsph relation obtained by correcting the K-band relation for the average B − K colour (i.e., CMR is not considered) with a BHMF derived with the B-band relation by MH03. In (4), the largest and smallest SMBH number densities are taken from the BHMFs in Figure 6. In addition to these, ∆ log φM due to the errors of characteristic luminosity and normal- ization in a COMBO-17 LF is shown for comparison in the column (7). The errors in the LF at z = 0.65 are adopted to calculate ∆ log φM here, but the errors in a BHMF are similar if LFs at other redshifts are used (see the lower panel of Figure 1). Figure 8. Top panel: The MBH − Lsph relations at z = 0.55 are calculated considering passive evolution of Lsph and are com- pared with the relation at z = 0 by MH03 (black solid line). We consider three models for passive luminosity evolution of old stel- lar population: zf = 10 (dashed line), 4 (grey solid line) and 3 (dotted line). Bottom panel: BHMFs at z = 0 and 0.55 derived using the above relations are indicated with the same line styles. Note that all the BHMFs (including that at z = 0) are calculated from the COMBO-17 LF at z = 0.55 (not iLF). e-folding time is 1 Gyr. In this case, the luminosity evolution in the B band is described as MB(z) = MB(z = 0) − Qz and Q = 1.4. The Q value can be estimated using other models of stellar populations and its dependency on choice of parameters may need to be treated as uncertainty of BHMF at high redshift. Since formation redshift (zf ) of a stellar population is probably the most important parameter upon which the Q value largely depends, we consider two other cases: zf = 10 and 3. The Q value is estimated to be 1.2 and 1.6, respectively. Using these Q values including 1.4 for zf = 4, we derive passively evolved B-band MBH − Lsph relations 8 N. Tamura, K. Ohta, & Y. Ueda log MBH ∆ log φM (1) (2) (3) (4) (5) (6) (7) 7.5 8.0 8.5 9.0 9.5 0.08 0.04 0.10 0.37 0.80 0.02 0.02 0.01 0.14 0.45 0.02 0.07 0.27 0.57 0.99 0.11 0.26 0.37 0.50 0.96 0.02 0.09 0.27 0.53 0.88 0.02 0.04 0.15 0.32 0.54 0.16 0.20 0.28 0.41 0.58 Table 1. Summary of uncertainties in a BHMF at the five BH masses. Each number shows the possible range of logarithmic SMBH mass density (∆ log φM ) at a given MBH caused by each source of uncertainty: (1) the 1σ fitting error of the B-band MBH−Lsph relation (see the middle panel of Figure 4), (2) the un- certainty of intrinsic scatter (0.3−0.4) in the MBH −Lsph relation (the bottom panel of Figure 4), (3) choice of filter bands (B or K) of MBH − Lsph relation (Figure 5), (4) choice of the MBH − Lsph relations by different authors (Figure 6), (5) B/T (0.5 − 0.9; Figure 7), (6) choice of passive evolution models (zf = 3 − 10; ∆ log φM is calculated at z = 1.05 (Figure 9)), and (7) errors of characteristic luminosity and normalization in a COMBO-17 LF (the errors of the LF at z = 0.65 are considered here; see the lower panel of Figure 1). In calculating ∆ log φM for (3), we compare a BHMF computed with the B-band MBH − Lsph relation obtained by correcting the K-band relation for the average B − K colour (i.e., CMR is not considered) with a BHMF derived with the B- band relation by MH03. In (4), the largest and smallest SMBH number densities are taken at each MBH. This table demonstrates that most of the uncertainties are negligible at MBH 6 108M⊙, while they are significant in the range of MBH > 108M⊙. The uncertainty related to the photometric band selection and colour correction and that related to selection of the MBH − Lsph relations by different studies are the most serious and they amount to an order of magnitude at the massive end. Those due to the fitting error of the B-band MBH − Lsph relation and the possible uncertainty in B/T also appear to be substantial. 3 BHMF FROM AGN LF In this section, we investigate the cosmological evolution of BHMFs derived from AGN LFs (AGN-BHMFs hereafter). The results will be compared with those from COMBO-17 LFs (spheroid-BHMFs hereafter) in § 4. By assuming that only mass accretion grows the central SMBH at a galactic centre and galaxy mergers are not important in its growth history, the time evolution of a BHMF φM(MBH, t) can be described by the continuity equation: ∂φM(MBH, t) ∂t + ∂ ∂MBH [φM(MBH, t)h M (MBH, t)i] = 0, (2) where h M (MBH, t)i represents the mean mass accretion rate at a given SMBH mass MBH and at a cosmic time t. Fur- thermore, if we assume a constant radiative efficiency ǫ (≡ L/ M c2, where L and M is the bolometric luminosity and the mass accretion rate, respectively) and a constant Eddington ratio λ (≡ L/LEdd, where LEdd is the Eddington luminosity) for all the AGNs, the second term of the above equation can be simply related to a (bolometric) luminosity function of AGNs. This finally reduces the continuity equa- tion to: ∂φM (MBH, t) ∂t = − (1 − ǫ)λ2c2 ǫt2 Edd ln 10 (cid:20) ∂ψ(L, t) ∂L (cid:21)L=λMBHc2/tEdd , (3) where tEdd is the Eddington time and ψ(L, t) is the AGN LF (for details, see Marconi et al. 2004). We note that ψ(L, t) is the number of AGNs per d log L, while φM (MBH, t) is the number of SMBHs per dMBH. Hence, once the form of an AGN LF is known as a function of redshift, one can integrate this equation to obtain BHMFs at any redshifts starting from the initial condition, either in time decreasing order (from a high redshift to z = 0) or the inverse (from z = 0 to higher redshifts). Following the procedure adopted by Marconi et al. (2004), here we derive AGN-BHMFs at intermediate red- shifts starting from a BHMF at z = 3 as the initial condition (it is assumed that all the SMBHs at z = 3 were shining as AGNs). In the calculation, we use the hard X-ray AGN LF (HXLF) by Ueda et al. (2003, U03 hereafter), which is de- scribed by a luminosity-dependent density evolution model (LDDE model; see their § 5.2 for details). To take into ac- count the contribution of "Compton-thick" AGNs to the total mass accretion rate, we multiply a correction factor of 1.6 independently of the AGN luminosity. The luminosity- dependent bolometric correction described in Marconi et al. (2004) is adopted. The Eddington ratio and the radiative ef- ficiency are assumed to be constant, λ = 1.0 and ǫ = 0.1, re- spectively, also based on the study by Marconi et al. (2004). The results are shown in Figure 10, indicating that while the BHMFs in the range of MBH > 108M⊙ hardly change out to z ∼ 1, they exhibit a clear redshift evolution at MBH 6 108M⊙. That is, almost all SMBHs with a mass larger than 108M⊙ formed at z & 1, while lighter SMBHs grow later, suggesting a downsizing of SMBH evolution. It needs to be pointed out that there are several uncer- tainties in the AGN-BHMFs thus far derived as follows: (i) There are ranges of values in Eddington ratio and ra- diative efficiency which give a reasonable fit of an AGN- BHMF to the spheroid-BHMF at z = 0. The χ2 distribution studied by Marconi et al. (2004) suggests a possible range of λ = 0.1 − 2.0 and ǫ = 0.04 − 0.15 within 1 σ uncer- tainty. In Figure 11, we exemplify AGN-BHMFs at z = 0 and 0.65 for five sets of λ and ǫ; either of λ or ǫ is fixed to the adopted value (λ = 1.0 or ǫ = 0.1) and the largest or smallest value within the uncertainty is chosen for the other parameter. This plot indicates that, as expected from equa- tion (3), the normalization of an AGN-BHMF is altered and the AGN-BHMF is shifted along the MBH axis by changing λ, while only the normalization is affected by changing ǫ. In the following analysis, AGN-BHMFs will be calculated for a number of pairs of λ and ǫ on the 1 σ contour provided by Marconi et al. (2004; see their Figure 7) and the envelope of these AGN-BHMFs that gives a possible range of SMBH density at a given MBH will be adopted as uncertainty of AGN-BHMF. (ii) The assumption of constant Eddington ratio and ra- diative efficiency for all the AGNs is perhaps too simple. Al- though Marconi et al. (2004) claim that the local spheroid- BHMF can be well reproduced by the AGN-BHMF with a constant λ ≃ 1.0 and ǫ ≃ 0.1, the solution only gives a suffi- cient condition to the limited constraints at z = 0. Further- more, Heckman et al. (2004) claim that AGNs have various Eddington ratios and the ratios seem to depend on MBH Supermassive black hole mass functions at intermediate redshifts 9 Figure 10. AGN-BHMF at a redshift of 0, 0.4, 0.8, 1.2 and 2.0 is plotted with open circles, solid circles, triangles, squares, and crosses, respectively. These AGN-BHMFs are calculated with the continuity equation and the HXLFs by U03 (see text for details). Figure 11. AGN-BHMFs at z = 0 (upper panel) and those at z = 0.65 (lower panel) for five sets of λ and ǫ as shown in the upper panel. Kawaguchi et al. (2004) also propose that super-Eddington accretion is essential for a major growth of SMBH. (iii) There is ambiguity in the continuity equation itself; if a merging process should be added as a source term in the equation, a resultant BHMF would change. (iv) Although the HXLF by U03 accounts for all the Compton-thin AGNs including obscured AGNs, the uncer- tainties in the estimate of Compton-thick AGNs directly af- fect the resulting BHMFs. Quantifying all of these uncertainties but (1) requires substantial theoretical works and/or new observational data and is beyond the scope of this paper, but one must keep them in mind. Figure 12. AGN-BHMFs at a redshift of 0.1, 0.6, and 1.2 ob- tained in this study by integrating the continuity equation start- ing from z = 3 are plotted with solid circles in the top, middle and bottom panel, respectively. The error bars represent the up- per and lower bounds of SMBH density at a given MBH that are calculated from AGN-BHMFs for the values of λ and ǫ within the 1 σ uncertainty. Asterisks show AGN-BHMFs obtained by Merloni (2004). It may be intriguing to compare these AGN-BHMFs with those recently obtained by Merloni (2004), where a new method is employed to investigate the redshift evolution. They introduce a conditional luminosity function (CLF), which is the number of active black holes per unit comoving volume per unit logarithm of radio (LR) and X-ray (LX) luminosity and is defined so that by integrating a CLF over the range of LR or LX , the radio LF (RLF) or X-ray LF (XLF) of AGNs is obtained, respectively. In order to relate LR and LX of AGN to MBH without any assumptions on ac- cretion rate, they use an empirical relation among LR, LX, and MBH (fundamental plane of black hole activity; Mer- loni, Heinz & Di Matteo 2003). A CLF also needs to satisfy a constraint that the number density of SMBHs with a cer- tain mass is obtained by integrating a CLF over the ranges of LR and LX which are determined by the fundamental 10 N. Tamura, K. Ohta, & Y. Ueda plane. Consequently, once a BHMF is obtained at a redshift of z, a CLF can be computed using RLF, XLF, and BHMF as constraints.3 Note that RLFs and XLFs from observa- tions are available up to high redshifts; the RLFs obtained by Willott et al. (2001) and the HXLF by U03 are used in Merloni (2004). Given a functional form of accretion rate4, a mean accretion rate can also be calculated as a function of MBH. Using a BHMF and a mean accretion rate at z, a BHMF at z + dz can be derived from the continuity equa- tion. Likewise, BHMFs at higher redshifts can successively be calculated and hence RLF, XLF, and BHMF at z = 0 are firstly needed. In Merloni (2004), the local BHMF is ob- tained using LFs of galaxies with different morphologies by Marzke et al. (1994) and the empirical relationships among spheroid luminosity, velocity dispersion, and MBH. Figure 12 shows the comparison of AGN-BHMFs at redshifts of 0.1, 0.6, and 1.2 calculated in the two differ- ent methods: (I) BHMFs derived by integrating the conti- nuity equation from z = 3 to lower redshifts using the U03 HXLF with λ = 1.0 and ǫ = 0.1. The error bars represent the upper and lower bounds of SMBH density at a given MBH that is calculated from AGN-BHMFs for a number of pairs of λ and ǫ on the 1 σ contour provided by Marconi et al. (2004). (II) Those obtained by Merloni (2004), which are integrated from z = 0 to higher redshifts using the CLF with the fundamental-plane relation. This comparison indi- cates that while the agreement of the AGN-BHMFs is good in the massive end, the discrepancy at MBH 6 107.5M⊙ is substantial at all redshifts. There could be several rea- sons for this disagreement. One possibility may be related to the choice of an initial spheroid-BHMF; we note that the BHMF at z = 0.1 in Merloni (2004) has a normalization ∼ 10 times smaller at MBH ∼ 106M⊙ than the local BHMF independently estimated by Marconi et al. (2004, see their Figure 2). Another reason could be the different estimate for the mean mass accretion rate. In fact, we find that the second term of the equation (2) averaged between z = 0.9 and z = 0.1 calculated in method I is significantly larger than that in method II in the range of MBH 6 107.5M⊙. We just point out the facts in this paper and leave further dis- cussions for future studies. In the next section, we adopt the AGN-BHMFs calculated by method I for comparison with the spheroid-BHMFs. 4 DISCUSSIONS 4.1 Comparison of Spheroid-BHMFs with AGN-BHMFs In Figure 13, spheroid-BHMFs are compared with AGN- BHMFs up to z ∼ 1. The spheroid-BHMFs transformed from the COMBO-17 LFs at redshifts of 0., 0.25, 0.45, 0.65, 0.85, and 1.05 are indicated with shaded regions (the BHMF at z = 0 was derived from the LF of the red sequence galax- ies in the SDSS EDR; see Appendix in Bell et al. 2004b). 3 In this calculation, the fundamental plane of black holes is as- sumed to be independent of redshift. 4 In Merloni εacc Mc2/LEdd is adopted (εacc is accretion efficiency). (2004), LX/LEdd = f (M, m) where m ≡ The widths of the shaded regions are determined by consid- ering not only the errors of M ∗ B and φ∗ but also the uncer- tainty associated with band transformation between B and K of the MBH − Lsph relation, which is the most significant uncertainty among those investigated (see § 2.3). The upper bound of the shaded region is the BHMF derived by applying the B-band MBH −Lsph relation by MH03 to an LF with the largest values of characteristic luminosity and normalization within the errors. To determine the lower bound, a B-band MBH − Lsph relation calculated by correcting the K-band relation for the average B − K colour of early-type galaxies is applied to an LF with the smallest values of characteristic luminosity and normalization within the errors. The trans- formation of the K-band relation to B band is performed only at z = 0 and the B-band relations at high redshifts are then obtained by correcting it for passive luminosity evolu- tion in the B band. The spheroid-BHMF is depicted with shaded region down to the mass corresponding to the lowest luminosity among the data points in the LF. At the masses lower than this cutoff, the upper- and lower-bound BHMFs are indicated by dashed lines. The characteristic MBH corre- sponding to L∗ B at each redshift is indicated by arrow. Two arrows are shown in each panel; the one at the more mas- sive end shows the characteristic MBH of the upper-bound BHMF, and the other is of the lower-bound BHMF. The solid curve, which goes through the middle of the upper- and lower-bound BHMFs at z = 0.25, indicates the BHMF at z = 0.255 calculated with a B-band MBH − Lsph rela- tion converted from the K-band relation by correcting for T-CMR. This BHMF is plotted in all the panels as a fidu- cial of comparison. The AGN-BHMFs calculated with the continuity equation and the HXLFs by U03 at the same redshifts as the spheroid-BHMFs are overplotted with open circles. The error bars indicate the upper and lower bounds of SMBH density allowing for the 1 σ uncertainty of λ and ǫ. From Figure 13, it is suggested that at MBH > 108M⊙, the spheroid-BHMFs6 are broadly consistent with the AGN- BHMFs out to z ∼ 1. This agreement between the spheroid- BHMFs and the AGN-BHMFs appears to support that most of the SMBHs are hosted by massive spheroids already at z ∼ 1 and they evolve without significant mass growth since then. The discrepancy at MBH 6 107.5M⊙ between the spheroid-BHMFs and the AGN-BHMFs is presumed to be due at least partly to the fact that small bulges in late-type galaxies tend to be excluded in selecting the red sequence galaxies and thus their contributions are not expected to be included in the COMBO-17 LFs or the spheroid-BHMFs. In fact, galaxy LFs are not well constrained down to such low luminosities and future observations therefore need to 5 We adopt the BHMF not at z = 0 but at z = 0.25 just for consistency; the LFs at z > 0.25 are calculated with the data from the COMBO-17 survey, while the LF at z = 0 is from the SDSS data. 6 In Figure 13, the spheroid-BHMF at z = 0.25 appears to exceed that at z = 0. This is because of the large increase of character- istic luminosity in the COMBO-17 LF from z = 0 to 0.25. This increase is significantly larger than that predicted for passive evo- lution, while the rate of luminosity evolution at z > 0.25 is fully consistent with passive evolution out to z ∼ 1. The origin of the large luminosity increase at the low redshift is currently unknown. Supermassive black hole mass functions at intermediate redshifts 11 Figure 13. The spheroid-BHMFs transformed from the COMBO-17 LFs at redshifts of 0, 0.25, 0.45, 0.65, 0.85, and 1.05 are indicated with shaded regions, whose widths are determined by considering not only the errors of M ∗ B and φ∗ but also the uncertainties associated with band transformation in the derivation of a BHMF (see text for details). The spheroid-BHMF at z = 0 was derived from the LF of the red sequence galaxies in the SDSS EDR (see Appendix of Bell et al. 2004b). The low mass cutoff of the shaded region corresponds to the lowest luminosity among the data points in the original LF. In the lower mass range than the cutoff, the upper- and lower-bound BHMFs are indicated by dashed lines. The characteristic MBH corresponding to M ∗ B at each redshift is indicated by arrow. Two arrows are shown in each panel; the one at the more massive end shows the characteristic MBH of the upper-bound BHMF, and the other is of the lower-bound BHMF. Solid curve indicates a BHMF at z = 0.25 calculated with a B-band MBH − Lsph relation converted from the K-band relation by correcting for T-CMR. This BHMF is displayed in all the panels as a fiducial of comparison. The AGN-BHMFs calculated with the continuity equation and the HXLFs by U03 at the same redshifts as those of the spheroid-BHMFs are overplotted with open circles. The error bars indicate the upper and lower bounds of SMBH density allowing for the 1 σ uncertainty of λ and ǫ as in Figure 12. 12 N. Tamura, K. Ohta, & Y. Ueda be awaited for any further discussions on the discrepancies in the light end of BHMF. It is interesting to point out that while the AGN-BHMFs at MBH > 108M⊙ do not signif- icantly evolve out to z ∼ 1, the spheroid-BHMFs exhibit a slight redshift evolution (see also Figure 1 and Figure 10). We note that the uncertainties of the spheroid-BHMFs considered in Figure 13 are larger than those in Figure 1, where the evolution in spheroid-BHMF may look clearer. One possible reason for the difference in evolution between the spheroid-BHMFs and the AGN-BHMFs is a selection effect on the red sequence galaxy LFs; if there are more blue spheroids with on-going star formation and/or young stellar population towards z = 1 then their contribution is more likely to be missed from the red sequence galaxy LF and the spheroid-BHMF at higher redshifts. In fact, as men- tioned earlier, the BHMFs obtained from morphologically selected early-type galaxy LFs tend to exceed those from the COMBO-17 LFs at z ∼ 1 (Figures 2 and 3). 4.2 Does the correlation between MBH and host spheroid mass evolve with redshift? An alternative interpretation of the possible difference in evolution between the spheroid-BHMFs and the AGN- BHMFs may be a difference at high redshift between the actual MBH − Lsph relation and our assumption. In other words, there may be an evolution of the MBH − Lsph rela- tion other than the passive luminosity evolution of spheroid. In order to see whether this can be the case, it is worth examining how a BHMF can be affected by changing the MBH − Lsph relation. Here, we consider a simple case where, in the MBH−Lsph relation (log MBH = p log Lsph + q), the coefficient of p or q varies. Figures 8 and 9 demonstrate the effect on BHMF of an evolution of q; a BHMF is shifted mostly in parallel to the MBH axis. The effect of changing p on a BHMF is sim- ilar, although it can be a modification in shape of BHMF rather than a lateral shift. Therefore, changing neither p nor q moves the BHMF along the axis of the number den- sity of SMBH. On the other hand, the redshift evolution of the actual spheroid-BHMFs seems to be dominated by that along the number density axis, suggesting that it is diffi- cult to explain the slight difference in evolution between the spheroid-BHMF and the AGN-BHMF by changing p and q with redshift. Recently, a possible offset from the local relationship between MBH and central velocity dispersion (σ0) has been found at z ∼ 0.37 by investigating spectra of the central regions of galaxies hosting type 1 AGNs (Treu, Malkan & Blandford 2004, T04 hereafter). It is interesting to see a spheroid-BHMF at this redshift derived with the offset found by T04. In order to apply the relation at z = 0.37 to a COMBO-17 LF at z = 0.35, the MBH − σ0 relation needs to be converted to a MBH − Lsph relation. One possible way is to estimate spheroid luminosities of the galaxies observed by T04 from their central velocity dispersions7 with the Faber- Jackson relation at z ∼ 0.4 in the rest-frame B band (Ziegler 7 According to T04, the velocity dispersions within an aperture used by T04 (σap) are converted to the central velocity dispersions (σ0) as σ0 = 1.1σap. Figure 14. Upper panel: The MBH − Lsph relations at z = 0 (MH03; black line) and 0.35 (grey line) are shown. For the relation at z = 0.35, passive evolution of Lsph is taken into account. The data points indicate the galaxies observed by T04 (see text for details). Lower panel: Shaded region indicates a spheroid-BHMF and its uncertainty obtained from the COMBO-17 LF at z = 0.35 and the MBH − Lsph relation at z = 0, which seems to be followed by the T04 galaxies. Solid lines show a BHMF at z = 0.25 obtained with a MBH − Lsph relation considering a passive luminosity evolution between z = 0 and 0.25. The upper and lower bounds are determined in the same way as those in Figure 13. The AGN-BHMF at z = 0.35 is overplotted with open circles. The error bars are calculated in the same way as those in Figure 12. et al. 2005) and plot them against MBH. The results are shown with solid circles in the upper panel of Figure 14. Although the sample size is small and there is a substantial scatter, the distribution of the data points suggests that the actual MBH − Lsph relation followed by these spheroids lies closer to the local relationship than that allowing for a pas- sive evolution of Lsph between z = 0 and 0.37. In the lower panel of Figure 14, a BHMF derived from the COMBO-17 LF at z = 0.35 using the MBH−Lsph relation at z = 0, which seems to be followed by the T04 galaxies, is indicated with shaded region. This BHMF is compared with that at a lower redshift (z = 0.25) derived with a MBH − Lsph relation con- sidering a passive luminosity evolution between z = 0 and 0.25. The latter BHMF is indicated with solid lines. The un- certainties of these spheroid-BHMFs are indicated with the width of the shaded region or the separation of the lines; the upper and lower bounds are defined in the same way as those in Figure 13. The AGN-BHMF at z = 0.35 is overplot- ted with open circles. The error bars are calculated in the same way as those in Figure 12. Although the uncertainties of the BHMFs are large, this comparison suggests that the BHMF obtained with the MBH − Lsph relation followed by the T04 galaxies is slightly more biased to the massive end. Supermassive black hole mass functions at intermediate redshifts 13 Heckman T. M., Kauffmann G., Brinchmann J., Charlot S., Tremonti C., White S. D. M., 2004, ApJ, 613, 109 Im M., et al., 2002, ApJ, 571, 136 Kawaguchi T., Aoki K., Ohta K., Collin S., 2004, A&A, 420, L23 Kochanek C. S., et al., 2001, ApJ, 560, 566 Kodama T., Arimoto N., 1997, A&A, 320, 41 Kodama T., Arimoto N., Barger A. J., Arag´on-Salamanca A., 1998, A&A, 334, 99 Magorrian J., et al., 1998, AJ, 115, 2285 Marconi A., Hunt L. K., 2003, ApJ, 589, L21 (MH03) Marconi A., Risaliti G., Gilli R., Hunt L. K., Maiolino R., Salvati M., 2004, MNRAS, 351, 169 Marzke R. O., Geller M. J., Huchra J. P., Corwin H. G., 1994, AJ, 108, 437 McLure R. J., Dunlop J. S., 2002, MNRAS, 331, 795 (MD02) McLure R. J., Dunlop J. S., 2004, MNRAS, 352, 1390 Merloni A., 2004, MNRAS, 353, 1035 Merloni A., Heinz S., Di Matteo T., 2003, MNRAS, 345, 1057 Merritt D., Ferrarese L., 2001, MNRAS, 320, L30 (MF01) Peletier R. F., Davies R. L., Illingworth G. D., Davis L. E., Cawson M., 1990, AJ, 100, 1091 Robertson B., Hernquist L., Cox T. J., Di Matteo T., Hop- kins P. F., Martini P., Springel V., 2005, ApJ, submitted (astro-ph/0506038) Schweizer F., Seitzer P., 1992, AJ, 104, 1039 Shankar F., Salucci P., Granato G. L., De Zotti G., Danese L., MNRAS, 354, 1020 Shields G. A., Gebhardt K., Salviander S., Wills B. J., Xie B., Brotherton M. S., Yuan J., Dietrich M., 2003, ApJ, 583, 124 Small T. A., Blandford R. D., 1992, MNRAS, 259, 725 Stoughton, C., et al., 2002, AJ, 123, 485 Tamura N., Ohta K., 2003, AJ, 126, 596 Taniguchi Y., et al., 2005, in the Proceedings of the 6th East Asian Meeting on Astronomy, JKAS, 39, in press (astro-ph/0503645) Terlevich A. I., Caldwell N., Bower, R. G., 2001, MNRAS, 326, 1547 Treu T., Malkan M. A., Blandford R. D., 2004, ApJ Letters, 615, L97 (T04) Ueda Y., Akiyama M., Ohta K., Miyaji T., 2003, ApJ, 598, 886 (U03) Willott C. J., Rawlings S., Blundell K. M., Lacy M., Eales S. A., 2001, MNRAS, 322, 536 Wolf C., Meisenheimer K., Rix H.-W., Borch A., Dye S., Kleinheinrich M., 2003, A&A, 401, 73 Ziegler B. L., Thomas D., Bohm A., Bender R., Fritz A., Maraston C., 2005, A&A, 433, 519 This trend could be seen more clearly when the BHMF is compared with that at z = 0. This apparently suggests a growth of SMBH towards higher redshift, which is unlikely in practice (see also Robertson et al. 2005). More galaxies need to be investigated to examine the correlation between MBH and σe or Lsph at high redshift. 4.3 Future perspective If much fainter end of early-type galaxy LF was determined from observations, BHMF could be probed further down to the low mass end (MBH ≪ 108M⊙). Since a redshift evolution of AGN-BHMF is suggested to be fast at inter- mediate redshifts in this mass range (Figure 10; see also Merloni 2004), it would be interesting to look at the coun- terpart in spheroid-BHMFs in order to put much stronger constraints on the co-evolution of AGN and spheroid, par- ticularly down-sizing effects of their evolutions. Currently, galaxy LFs are not well constrained down to such low lu- minosities, especially at cosmological distances. Even at low redshifts, the faint end slope of an LF tends to be fixed to a certain value in fitting a Schechter function to the data and deriving L∗ and φ∗. Much deeper data (e.g., ∼ 2 mag deeper than the COMBO-17 limit) are essential to address the faint end of LF out to z ∼ 1. Keeping a survey area similar to or even wider than COMBO-17 is the key to de- riving reliable LFs and to estimate cosmic variance. Multi- band photometry would be required to obtain photometric redshifts in good accuracy down to the faint end. In addi- tion, high spatial resolution images would enable one to di- rectly measure spheroid luminosities of galaxies. Although it is highly expensive to take data sets satisfying all these re- quirements, one promising candidate for this challenge is the COSMOS survey: a 2 square degree field is surveyed with the HST/ACS in the I814 band down to 27.8 mag (5σ) in the AB magnitude (cf. galaxies with R 6 24 mag are used to study LFs in COMBO-17) and also with Subaru/Suprime- Cam in the BV r′i′z′ bands (Taniguchi et al. 2005). ACKNOWLEDGMENTS We are grateful to the anonymous referee for helpful com- ments to improve this paper. This research was partly sup- ported by a Grant-in-Aid for Scientific Research from Japan Society for the Promotion of Science (17540216). REFERENCES Aller M. C., Richstone D., 2002, AJ, 124, 3035 Bell E. F., et al., 2004a, ApJ, 600, L1 Bell E. F., et al., 2004b, ApJ, 608, 752 Bower R. G., Lucey J. R., Ellis R. S., 1992, MNRAS, 254, 601 Cavaliere A., Morrison P., Wood K., 1971, ApJ, 170, 223 Cross N. J. G., et al., 2004, AJ, 128, 1990 Ferrarese L., Merritt D., 2000, ApJ, 539, L9 (FM00) Fioc M., Rocca-Volmerange B., 1997, A&A, 326, 950 Gebhardt, K., et al., 2000, ApJ, 539, L13 Girardi M., Mardirossian F., Marinoni C., Mezzetti M., Rigoni E., 2003, A&A, 410, 461
astro-ph/0005414
1
0005
2000-05-19T22:01:33
The Discovery of a Luminous z=5.80 Quasar from the Sloan Digital Sky Survey
[ "astro-ph" ]
We present observations of SDSSp J104433.04--012502.2, a luminous quasar at z=5.80 discovered from Sloan Digital Sky Survey (SDSS) multicolor imaging data. This object was selected as an i'-band dropout object, with i*=21.8 +/- 0.2, z*=19.2 +/- 0.1. It has an absolute magnitude M1450 = -27.2 (H_0 =50 km/s/Mpc, q0 = 0.5). The spectrum shows a strong and broad Ly alpha emission line, strong Ly alpha forest absorption lines with a mean continuum decrement D_A = 0.91, and a Lyman Limit System at z=5.72. The spectrum also shows strong OI and SiIV emission lines similar to those of quasars at z<= 5, suggesting that these metals were produced at redshift beyond six. The lack of a Gunn-Peterson trough in the spectrum indicates that the universe is already highly ionized at z ~ 5.8. Using a high-resolution spectrum in the Ly alpha forest region, we place a conservative upper limit of the optical depth due to the Gunn-Peterson effect of tau < 0.5 in regions of minimum absorption. The Ly alpha forest absorption in this object is much stronger than that in quasars at z<= 5. The object is unresolved in a deep image with excellent seeing, implying that it is unlensed. The black hole mass of this quasar is ~3 x 10^9 M_solar if we assume that it is radiating at the Eddington luminosity and no lensing amplification, implying that it resides in a very massive dark matter halo. The discovery of one quasar at M_1450 < -27 in a survey area of 600 deg^2 is consistent with an extrapolation of the observed luminosity function at lower redshift. The abundance and evolution of such quasars can provide sensitive tests of models of quasar and galaxy formation.
astro-ph
astro-ph
The Discovery of a Luminous z = 5.80 Quasar from the Sloan Digital Sky Survey1 Xiaohui Fan2 , Richard L. White3 , Marc Davis4 , Robert H. Becker5,6 , Michael A. Strauss2 , Zoltan Haiman2 , Donald P. Schneider7 , Michael D. Gregg5,6 , James E. Gunn2 , Gillian R. Knapp2 , Robert H. Lupton2 , John E. Anderson, Jr.8 , Scott F. Anderson9 , James Annis8 , Neta A. Bahcall2 , William N. Boroski8 , Robert J. Brunner10 , Bing Chen11 , Andrew J. Connolly12 , Istvan Csabai11 , Mamoru Doi13 , Masataka Fukugita14,15 , G. S. Hennessy16 , Robert B. Hindsley17 , Takashi Ichikawa18 , Zeljko Ivezi´c2 , Avery Meiksin19 , Timothy A. McKay20 , Jeffrey A. Munn21 , Heidi J. Newberg22 , Robert Nichol23 , Sadanori Okamura13 , Jeffrey R. Pier21 , Maki Sekiguchi14 , Kazuhiro Shimasaku13 , Alexander S. Szalay11 , Gyula P. Szokoly24 , Aniruddha R. Thakar11 , Michael S. Vogeley25 , Donald G. York26 0 0 0 2 y a M 9 1 1 v 4 1 4 5 0 0 0 / h p - o r t s a : v i X r a – 2 – 1Based on observations obtained with the Sloan Digital Sky Survey, which is owned and operated by the Astrophysical Research Consortium, and at the W. M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California, and NASA, and was made possible by the generous financial support of the W. M. Keck Foundation. 2Princeton University Observatory, Princeton, NJ 08544 3Space Telescope Science Institute, Baltimore, MD 21 218 4Department of Astronomy, University of California, Berkeley, CA 94720-3411 5Physics Department, University of California, Davis, CA 95616 6 IGPP/Lawrence Livermore National Laboratory, Livermore, CA 95616 7Department of Astronomy and Astrophysics, The Pennsylvania State University, University Park, PA 16802 8Fermi National Accelerator Laboratory, P.O. Box 500, Batavia, IL 60510 9University of Washington, Department of Astronomy, Box 351580, Seattle, WA 98195 10 Department of Astronomy, California Institute of Technology, Pasadena, CA 91125 11 Department of Physics and Astronomy, The Johns Hopkins University, 3701 San Martin Drive, Baltimore, MD 21218, USA 12 Department of Physics and Astronomy, University of Pittsburgh, Pittsburgh, PA 15260 13Department of Astronomy and Research Center for the Early Universe, School of Science, University of Tokyo, Hongo, Bunkyo, Tokyo, 113-0033, Japan 14 Institute for Cosmic Ray Research, University of Tokyo, Midori, Tanashi, Tokyo 188-8502, Japan 15 Institute for Advanced Study, Olden Lane, Princeton, NJ 08540 16U.S. Naval Observatory, 3450 Massachusetts Ave., NW, Washington, DC 20392-5420 17Remote Sensing Division, Code 7215, Naval Research Laboratory, 4555 Overlook Ave. SW, Washington, DC 20375 18Astronomical Institute, Tohoku University, Aoba, Sendai 980-8578 Japan 19 Institute for Astronomy, University of Edinburgh, Edinburgh EH9 3HJ, UK 20University of Michigan, Department of Physics, 500 East University, Ann Arbor, MI 48109 21U.S. Naval Observatory, Flagstaff Station, P.O. Box 1149, Flagstaff, AZ 86002-1149 22Rensselaer Polytechnic Insitute, Dept. of Physics, Applied Physics, and Astronomy, Troy, NY 12180 23 Department of Physics, Carnegie Mellon University, Pittsburgh, PA 15213 24 Astrophysikalisches Institut Potsdam An der Sternwarte 16 D-14482 Potsdam, Germany 25Department of Physics, Drexel University, 3141 Chestnut St., Philadelphia, PA 19104 – 3 – ABSTRACT We present observations of SDSSp J104433.04–012502.2, a luminous quasar at z = 5.80 discovered from Sloan Digital Sky Survey (SDSS) multicolor imaging data. This ob ject was selected as an i′ -band dropout ob ject, with i∗ = 21.8 ± 0.2, z ∗ = 19.2 ± 0.1. It has an absolute magnitude M1450 = −27.2 (H0 = 50 km s−1 Mpc−1 , q0 = 0.5). The spectrum shows a strong and broad Lyα emission line, strong Lyα forest absorption lines with a mean continuum decrement DA = 0.91, and a Lyman Limit System at z = 5.72. The spectrum also shows strong OI and SiIV emission lines similar to those of quasars at z ∼< 5, suggesting that these metals were produced at redshift beyond six. The lack of a Gunn-Peterson trough in the spectrum indicates that the universe is already highly ionized at z ∼ 5.8. Using a high-resolution spectrum in the Lyα forest region, we place a conservative upper limit of the optical depth due to the Gunn-Peterson effect of τ < 0.5 in regions of minimum absorption. The Lyα forest absorption in this ob ject is much stronger than that in quasars at z ∼< 5. The ob ject is unresolved in a deep image with excellent seeing, implying that it is unlensed. The black hole mass of this quasar is ∼ 3 × 109M⊙ if we assume that it is radiating at the Eddington luminosity and no lensing amplification, implying that it resides in a very massive dark matter halo. The discovery of one quasar at M1450 < −27 in a survey area of 600 deg2 is consistent with an extrapolation of the observed luminosity function at lower redshift. The abundance and evolution of such quasars can provide sensitive tests of models of quasar and galaxy formation. 1. Introduction At what epoch did the first generation of galaxies and quasars form? How was the universe re-ionized, ending the “dark ages” (Rees 1998)? These fundamental questions can only be answered with studies of high-redshift ob jects. The last few years have witnessed the first direct observations of galaxies at redshift higher than five (Dey et al. 1998, Weymann et al. 1998, Spinrad et al. 1998, Chen, Lanzetta & Pascarelle 1999, van Breugel et al. 1999, Hu et al. 1999, see also the review by Stern & Spinrad 1999), while detailed studies of the ensemble properties and large scale distribution of galaxies at z ∼ 4 have begun (Steidel et al. 1998, 1999). Several quasars have been found at z ∼> 5 (Fan et al. 1999, 2000a, Zheng et al. 2000), including a low-luminosity quasar at z = 5.50 (Stern et al. 2000). 26University of Chicago, Astronomy & Astrophysics Center, 5640 S. Ellis Ave., Chicago, IL 60637 – 4 – Studies of high-redshift quasars provide important probes of this critical epoch in cosmic evolution. The lack of the Gunn-Peterson (1965) effect in the absorption spectrum of a z ∼ 5.0 quasar (Songaila et al. 1999) indicates that the universe is already highly ionized at that redshift. The exact epoch of re-ionization could be determined from the absorption spectra of quasars at even higher redshift (Miralda-Escud´e 1997, Haiman & Loeb 1999). The study of the luminosity function of high-redshift quasars will constrain models of quasar and galaxy evolution (Haiman & Loeb 1998, Haehnelt et al. 1999), and determine whether it was UV radiation from AGNs or from young massive stars that re-ionized the universe, ending the “dark ages” (Haiman & Loeb 1998). Measurements of the chemical abundance in the quasar environment will reveal the metal production process at the very early stage of galaxy evolution (Hammann & Ferland 1999). Finally, luminous high-redshift quasars represent high peaks of density fields, and may be the markers of large scale structure at these early epochs (Djorgovski et al. 1999, Haiman & Hui 2000, Martini & Weinberg 2000). The Sloan Digital Sky Survey (SDSS; York et al. 2000) is using a dedicated 2.5m telescope and a large format CCD camera (Gunn et al. 1998) at the Apache Point Observatory in New Mexico to obtain images in five broad bands (u′ , g ′ , r ′ , i′ and z ′ , centered at 3540, 4770, 6230, 7630 and 9130 A, respectively; Fukugita et al. 1996) over 10,000 deg2 of high Galactic latitude sky. The multicolor data from SDSS have proven to be very effective in selecting high-redshift quasars: more than 50 quasars at z > 3.5 have been discovered to date from about 600 deg2 of imaging data (Fan et al. 1999, 2000a, Schneider et al. 2000, Zheng et al. 2000). The inclusion of the reddest band, z ′ , in principle enables the detection of quasars up to z ∼ 6.5 in SDSS data. In this paper, we report the discovery of SDSSp J104433.04–012502.2 (the name reflecting its J2000 coordinates from the preliminary SDSS astrometry, accurate to ∼ 0.1′′ in each coordinate), a very luminous, “i′ -dropout” quasar at z = 5.80, selected by its very red i∗ − z ∗ color. In a Λ-dominated flat universe (H0 = 65 km s−1 Mpc−1 , Λ = 0.65 and Ω = 0.35, referred to as the Λ-model in this paper, Ostriker & Steinhardt 1995, Krauss & Turner 1995), z = 5.80 corresponds to an age of 0.9 Gyr in an universe 13.9 Gyr old, or a look-back time of 93.2% of the age of the universe. Similarly, the universe was 0.7 Gyr old at z = 5.80 in a universe 13.0 Gyr old at present for a model with Ω = 1 and H0 = 50 km s−1 Mpc−1 , which we refer to as the Ω = 1 model in this paper. We present the photometric observations and target selection in §2, and the spectroscopic observations in §3. In §4, we discuss the cosmological implications, including the constraints on the Gunn-Peterson effect, quasar evolution models, and black hole formation. – 5 – 2. Photometric Observation and Target Selection The ob ject SDSSp J104433.04–012502.2 (hereafter SDSS 1044–0125 for brevity) was selected from the SDSS imaging data based on its extremely red i∗ − z ∗ color. The photometry of this ob ject is summarized in Table 1. The photometric observations of this region were obtained by the SDSS imaging camera on 2000 March 4 during the SDSS commissioning phase. For off-equator scans, the telescope moves along a great circle, with the photometric camera drift-scanning at the sidereal rate. The effective exposure time is 54.1 seconds in each band. The seeing in the i′ and z ′ bands was about 1.9′′ . The photometric calibration is provided by an auxiliary 20-inch telescope at the same site (Uomoto et al. 2000, in preparation). The photometric zeropoint is accurate to about 7% in u′ and z ′ and 2% in g ′ , r ′ and i′ . Because the definition of the photometric system is not yet finalized, we quote measured magnitudes using asterisks (e.g., i∗ ) to represent preliminary photometry, while referring to the filters with primes (e.g., i′ ). SDSS 1044–0125 is undetected in u′ , g ′ and r ′ . The ob ject is detected at the 5-σ level in the i′ band. The z ′ detection is of very high significance, with i∗ − z ∗ = 2.58 ± 0.20. Data are quoted as asinh magnitudes (Lupton, Gunn & Szalay 1999) and are on the AB magnitude system (Fukugita et al. 1996). Finding charts for SDSS 1044–0125 in the i′ and z ′ bands are shown in Figure 1. As shown by Fan (1999), the colors of high-redshift quasars are strong functions of redshift in the SDSS filter system, as first the Lyα forest and then the Lyman Limit Systems move through the filter system. At z > 3.6, quasars become very red in g ∗ − r∗ while remaining blue in r∗ − i∗ , and can be readily distinguished from stars based on these colors. At z > 4.6, quasars become very red in r∗ − i∗ . They are easily distinguished from red stars in the r∗ − i∗ vs. i∗ − z ∗ color-color diagram (Fan et al. 1999, 2000a). Quasars at z ∼ 5.5 have very similar r∗ − i∗ and i∗ − z ∗ colors to those of late M stars. Finally, at z ∼> 5.7, the Lyα emission line begins to move out of the SDSS i′ filter. With a predicted i∗ − z ∗ ∼> 2, quasars at such redshifts become i′ -band dropout ob jects. However, unlike the lower-redshift g ′ and r ′ -dropout quasars, these quasars have only one measurable optical color, as they will be completely undetected in r ′ . The SDSS alone cannot provide a constraint on the quasar’s continuum shape redward of the Lyα emission. It is thus difficult to distinguish them from other classes of red ob jects without additional information, such as near-infrared photometry (see also Zheng et al. 2000) or detection in the radio or X-ray. In fact, the only other known class of stellar ob jects with such red colors (i∗ − z ∗ ∼> 2) are the extremely cool stars/substellar ob jects with spectral type L or T (Kirkpatrick et al. 1999, Strauss et al. 1999, Tsvetanov et al. 2000, Fan et al. 2000b, Leggett et al. 2000). Most of these ob jects are brown dwarfs with mass below the hydrogen burning limit. – 6 – Although L and T dwarfs are very rare on the sky (the density for L dwarfs is 1 per ∼ 15 deg2 for i∗ < 20, Fan et al. 2000b), they are still more numerous than are z > 5 quasars, and are the ma jor contaminants of searches for i′ -dropout quasars. Near infrared photometry can be used to separate high-redshift quasar candidates from these cool dwarfs: the continuum shape of quasars is relatively flat towards near-IR bands, while the flux of L dwarfs continues to rise sharply towards longer wavelengths. We have selected i′ -dropout candidates with i∗ − z ∗ ∼> 2 and z ∗ < 19.5 from about 600 deg2 of SDSS imaging data. These 600 deg2 overlap the publicly released area of the Two Micron All Sky Survey (2MASS). SDSS 1044–0125 is the only i′ -dropout source in the 2MASS covered area that is not detected at the 7-σ level in any band in the 2MASS Point Source Catalog. All other sources are detected in at least one band at more than 10 σ (e.g. Fan et al. 2000b, Leggett et al. 2000). Is the non-detection in 2MASS sufficient to rule out the possibility that SDSS 1044–0125 is a brown dwarf ? In this area of the sky, the 2MASS Point Source Catalog has 7-σ limiting magnitudes of roughly J = 16.7, H = 15.9 and Ks = 15.0. This indicates a 7-σ upper limit on the optical-IR colors of SDSS 1044–0125: z ∗ − J < 2.5, z ∗ − H < 3.3 and z ∗ − Ks < 4.2. There are three L dwarfs in Fan et al. (2000b) that have i∗ − z ∗ > 2: SDSS 0330–0025 (i∗ − z ∗ = 2.13, spectral type L2), SDSS 0539–0059 (i∗ − z ∗ = 2.31, L5) and SDSS 1326–0038 (i∗ − z ∗ = 2.61, L8). They all have z ∗ − J ∼> 2.7, z ∗ − H ∼> 3.6 and z ∗ − Ks ∼> 4.2 (see Tables 2 and 5 in Fan et al. 2000b). Although the current L dwarf sample is small, the i∗ − z ∗ and the z ∗ − J (or H , Ks ) colors seem to be reasonable indicators of the spectral type for L dwarfs (Kirkpatrick et al. 1999, Fan et al. 2000b) Therefore, if SDSS 1044–0125 (i∗ − z ∗ = 2.6) had been an L dwarf, we would expect it to have been detected by 2MASS with high significance. Thus the non-detection in the 2MASS passbands indicates that SDSS 1044–0125 is unlikely to be a L dwarf, but is rather a source with a much flatter IR continuum, such as a quasar at z > 5.6 or a compact galaxy at z > 1. A K ′ -band image of SDSS 1044-0125 was obtained on the night of 2000 April 17, using the Near Infrared Camera (NIRC, Matthews & Soifer 1994) on the Keck I telescope under photometric skies and good seeing. The observations consisted of a nine-point dither pattern, integrating for 3 × 10s coadds at each location. The data were flattened, sky-subtracted, shifted, and stacked using the DIMSUM package in IRAF. The images show SDSS 1044-0125 to be an unresolved point source with FWHM = 0.′′375. Photometry through a 2′′ radius aperture yields K ′ = 17.02 ± 0.04, referenced to the standards of Persson et al. (1998). There are no companions or associated structure within 20′′ of the quasar to a 3σ point source limiting magnitude of K ′ ≈ 22.3. – 7 – 3. Spectroscopy Low and high dispersion spectra of SDSS 1044–0125 were obtained with the Echelle Spectrograph and Imager (ESI; Epps & Miller 1998) on the Keck II telescope on the night of 2000 April 6. The night was photometric with 0.9′′ seeing. A 1200 second exposure was taken through a 0.7′′ slit in the low-dispersion mode of ESI. In this mode, a prism is used for dispersion. The spectrum covers 3900 A to about 10000 A. The dispersion varies roughly linearly with wavelength from 0.8 A/pixel at 3900 A to 10 A/pixel at 10000 A. In addition, two 1200 second high resolution spectra were taken through a 1.0′′ slit in the echellette mode of ESI. In this mode, the spectral range of 3900 A to 11000 A is covered in ten spectral orders with a constant dispersion of 11.4 km s−1 pixel−1 . Wavelength calibrations were performed with observations of Hg-Ne-Xe lamps in the low-dispersion prism mode, and a Cu-Ar lamp in the echellette mode. The spectrophotometric standard G191-B2B (Massey 1988, Massey & Gronwall 1990) was observed for flux calibration. All observations were carried out at the parallactic angle. The data were reduced with standard IRAF routines. Figure 2 shows the final spectrum combining the low and high dispersion observations (total exposure time of 3600s), over a wavelength range of 4500 - 10000 A, and binned to 4 A/pixel. The telluric absorption bands were removed using the standard star spectrum. The absolute flux scale of the spectrum is adjusted so that it reproduces the SDSS z ∗ magnitude. The signal-to-noise ratio at λ > 8000 A is 15 – 20 per pixel. The spectrum of SDSS 1044–0125 shows the unambiguous signature of a very high redshift quasar: the broad and strong Lyα+NV emission line at λ ∼ 8300 A with a sharp discontinuity to the blue side, due to the onset of very strong Lyα forest absorption. The flux level drops by a factor of ∼ 10 from the red side to the blue side of the Lyα+NV emission. A Lyβ+OVI emission line is detected at ∼ 7000 A, with an additional flux decrement due to the onset of Lyβ forest absorption lines. The spectrum shows no detectable flux at λ < 6100 A because of the presence of a Lyman Limit System (see §4). Redward of Lyα, two additional emission lines, OI+SiIIλ1302 and SiIV+OIV]λ1400 are also clearly visible. The synthetic i∗ − z ∗ color calculated from the spectrum in Figure 2 is 2.5, consistent with the SDSS measurement. The redshift determination of such an ob ject is not straightforward. The Lyα emission line is severely affected by the strong Lyα forest lines (as well as by possible internal absorption lines). For z ∼ 4 quasars, Schneider, Schmidt & Gunn (1991) show that the peak of the Lyα emission line is typically at rest-frame 1219 A, but with a large scatter. Another strong line, CIVλ 1549, is out of the range of our spectrum. OI+SiIIλ1302 and SiIV+OIV]λ1400 are relatively weak. We use the central wavelengths of these two blends from Francis et al. (1991). Gaussian fits of these two lines are listed in Table 2. Based on – 8 – these fits, we adopt a redshift of 5.80 ± 0.02 for SDSS 1044–0125. Ob jective measurement of the equivalent width of the Lyα line is difficult (see the discussion by Schneider, Schmidt & Gunn 1991). Using the continuum level determined from the red side of Lyα emission, we find the rest-frame equivalent width of Lyα +NV is ∼ 26 A. However, this number is clearly an underestimate due to the extreme Lyα absorption, maybe by a factor of two. Figure 2 indicates that the expected line center of Lyα at z = 5.80 is severely absorbed, and that the blue wing of the line has almost completely disappeared. The measurement of the Lyβ+OVI line is difficult as well. We smooth the spectrum of Figure 2 to estimate the continuum level, and find an equivalent width of ∼ 30A. This value is also highly uncertain due to the difficulty in defining a “continuum” in the Lyα forest region. It is interesting to note that the strengths of the OI+SiIIλ1302 and SiIV+OIV]λ1400 line blends are comparable to those at much lower redshift: the average rest frame equivalent widths of these two lines are 3.2 ± 0.4 A and 8.1 ± 0.6 A in the sample of 30 quasars at z ∼ 4 in Schneider, Schmidt & Gunn (1991), compared to 3.4 and 7.0 A in the case of SDSS 1044–0125. Previous studies have shown that quasar environments at z ∼ 4 have roughly solar or higher metallicities (Hammann & Ferland 1999). Although we cannot derive the metallicity of SDSS 1044–0125 based on these emission line measurements, the existence of strong lines suggests that in this system, the metallicity is already quite high at this redshift. Assuming the metals were produced in stellar evolution, the initial starburst and chemical enrichment of the quasar environment must have happened at a very early epoch. From the spectrum in Figure 2, we derive its continuum AB magnitude at rest frame 1280 A, AB1280 = 19.28, after correcting for interstellar extinction (E (B − V ) = 0.054, from the map of Schlegel, Finkbeiner & Davis 1998). In the Λ-model (see §1), SDSS 1044–0125 has an absolute magnitude M1280 = −27.41. Assuming a power law continuum with fν ∝ ν −0.5 we find M1450 = −27.50 and MB = −27.96. In the Ω = 1 model (§1) it has M1280 = −27.15, M1450 = −27.24, and MB = −27.70. In this cosmology, the nearby luminous quasar 3C273 has MB = −27.0. SDSS 1044–0125 is a very luminous quasar, about twice as luminous as 3C 273 (assuming that it is not amplified by lensing or beaming). SDSS 1044–0125 is detected neither in the FIRST radio survey (Becker, White & Helfand 1995) at the 1mJy level at 20cm wavelength, nor in the ROSAT All Sky Survey (Voges et al. 1999), implying a 3-σ upper limit of 3 × 10−13 ergs cm−2 s−1 in the 0.1 – 2.4 keV band. These result is not unexpected; only a few z > 4 quasars have observed X-ray or radio fluxes above this value (Kaspi, Brandt & Schneider 2000, Schmidt et al. 1995). A deep exposure with Chandra or XMM is needed to determine its X-ray properties. – 9 – 4. Discussion 4.1. Absorption Properties and Gunn-Peterson Effect The high luminosity of SDSS 1044–0125 makes it an ideal ob ject for high signal-to-noise ratio observations to study the intergalactic medium at high redshift. In order to detect continuum break caused by the Lyman Limit System, an edge filter with width of 40 A was convolved with the low resolution spectrum in Figure 2. A strong peak at 6131 A is detected in the convolved spectrum, indicating the existence of a Lyman Limit System at zLLS = 5.72. No flux is detected blueward of this break. A Lyman Limit System is usually detected within 0.1 of the emission redshift in essentially all quasars at z > 4 (Schneider, Schmidt & Gunn 1991, Storrie-Lombardi et al. 1996, Fan et al. 1999). The most striking feature of the spectrum of SDSS 1044–0125 is the very strong absorption caused by Lyα forest lines. However, the flux level in the Lyα forest region never reaches zero. It lacks the Gunn-Peterson (1965) trough that would exist in the spectrum of a quasar at redshift higher than the re-ionization redshift (Haiman & Loeb 1999), indicating that the intergalactic medium is already highly ionized at z ∼ 5.8. We estimate ν E, where f obs the average continuum decrements as: DA,B ≡ D1 − f obs and f con ν /f con are the ν ν observed and the unabsorbed continuum fluxes of the quasar, and DA and DB measure the decrements in the region between rest-frame Lyα and Lyβ (λ = 1050 − 1170 A) and between Lyβ and the Lyman Limit (λ = 920 − 1050 A), respectively (Oke & Korycansky 1982). The measurements of DA and DB require knowledge of the continuum shape redward of Lyα. However, with DA approaching unity, the effect of different slopes is quite small. Assuming a power law continuum ν α with α = −0.5, as indicated in Figure 3, we obtain DA = 0.91 and DB = 0.94. Using a slope of –1.0 only changes the DA and DB values to 0.92 and 0.96, respectively. We therefore adopt DA = 0.91 ± 0.02 and DB = 0.95 ± 0.02. These values are in close accordance with those from quasar RD J030117+002025 (z = 5.50, Stern et al. 2000) and from distant galaxies in the Hubble Deep Field (Weymann et al. 1998), and are much higher than that of the z = 5.00 quasar SDSSp J033829.31+002156.3 (DA = 0.75, Songaila et al. 1999), suggesting that the strong evolution of the strength of the Lyα forest at z > 5, N (z) ∝ (1 + z)2.3−2.75 , measured at redshifts below five continues to a redshift of nearly six. Assuming this number density evolution, Zuo (1993) and Fan (1999) show that at z ∼ 5.8, the expected average DA ranges from 0.8 to 0.9. We further derive an upper limit on the Gunn-Peterson optical depth following Songaila et al. (1999). Figure 3 shows the high-resolution echellette spectrum of SDSS 1044–0125 in the Lyα forest region (binned to 2 A/pixel). The continuum level is approximated by a ν −0.5 power law as above. Even the most transparent part of the forest does not return close – 10 – to the continuum level at this resolution. It is evident that the Lyα forest is much stronger in SDSS 1044–0125 than in SDSSp J033829.31+002156.3 (Figure 2 of Songaila et al. 1999), where a fraction of the forest has flux comparable to the extrapolated continuum. In the region between 7926 A and 7929 A, SDSS 1044–0125 has an optical depth τ = 0.35 ± 0.07, where the error bar only reflects the statistical noise in the spectrum (note the noise level indicated in the figure). This value changes to 0.40 and 0.31 for power law slopes of 0.0 and −1.0, respectively. Therefore, we adopt a conservative limit of τ < 0.5 at this redshift of 5.52. With higher resolution and signal-to-noise ratio, we might be able to select regions even less affected by the Lyα forest lines. Therefore, it is only an upper limit. For comparison, Songaila et al. derived τ < 0.1 for z = 4.72 with similar resolution. The Gunn-Peterson effect analysis above is based on an attempt to measure the amount of flux between Lyα forest lines (e.g., Giallongo et al. 1994). At redshift higher than five, even with a moderately high resolution spectrum, these forest lines overlap, making it impossible to find a truly “line-free” region. Modern hydrodynamic simulations and semi-analytic models show that under the influence of gravity, the intergalactic medium becomes clumpy, and the Gunn-Peterson optical depth should vary even in the lowest column density regions (e.g. Bi, Borner & Chu 1992, Miralda-Escud´e & Rees 1993, Cen et al. 1993, Hernquist et al. 1996). The minimum absorption regions in the forest merely represent regions that are most underdense in this fluctuating Gunn-Peterson effect. An accurate measurement of the Gunn-Peterson effect and the ionizing background from the high resolution spectrum of SDSS 1044–0125 requires detailed comparison with cosmological simulations; this is beyond the scope of the current paper. Figure 2 also shows the detection of an intervening MgII absorption system. The MgII doublet λ2796.4 + 2803.5 is detected at wavelengths 9166.7 A and 9190.3 A in the high-resolution spectrum; the redshift of this system is zabs = 2.278. The rest frame equivalent widths of the doublet lines are 2.43 and 1.90 A, respectively. This system is very similar to the one detected in SDSSp J033829.31+002156.3 (zabs = 2.304, Songaila et al. 1999). It is possible that SDSS 1044–0125 is amplified by lensing from this intervening absorber. However, we saw in § 2 that this ob ject is unresolved under 0.4′′ seeing in the K band. 4.2. Number Density of Very High Redshift Quasars The total area of SDSS imaging data that we have searched for high-redshift quasars thus far is of order 600 deg2 . All that satisfy z ∗ < 19.3 and i∗ − z ∗ > 2.2 in this 600 deg2 region have been observed spectroscopically. Only SDSS 1044–0125 is identified as – 11 – a high-redshift quasar; the remaining ob jects are L and T dwarfs. Using the luminosity function and redshift dependence of Schmidt, Schneider & Gunn (1995) (for the Ω = 1 model), extrapolating it to higher redshifts and assuming fν ∝ ν −0.5 , we predict that in a total area of 600 deg2 , for z > 5.65 (i′ -dropout ob jects ), there should be 1.5 quasars with M1450 < −27.0 and 1.1 quasars with M1450 < −27.2. For z > 5.8, the extrapolation predicts 1.1 and 0.8 quasars for M1450 < −27.0 and −27.2, respectively. This assumes that our selection efficiency is 100%. The Schmidt-Schneider-Gunn luminosity function is derived using ob jects with 2.7 < z < 4.7 and −27.5 < MB < −25.5. Although it is difficult to draw any reliable conclusion from the observation of a single high-redshift quasar, the discovery of SDSS 1044–0025 is consistent with the expectations from this rather large extrapolation from lower redshift results. Assuming that the same luminosity function holds at even higher redshift, the SDSS will be able to discover one quasar at z ∼> 6, z ∗ ∼< 19 in every 1500 deg2 of the survey. SDSS 1044-0125 is a very luminous quasar. Assuming that (1) its bolometric luminosity equals the Eddington luminosity, Lbol = LEdd = 1.5 × 1038(MBH /M⊙ ) erg s−1 ; (2) its intrinsic continuum spectrum is the same as the mean spectral template of Elvis et al. (1994), and (3) neither beaming nor lensing affects the observed flux, we find a black hole mass of MBH = 3.4 × 109M⊙ in the Λ–model, or MBH = 2.7 × 109M⊙ in the Ω = 1 model. In either case, the implied black hole mass is quite large, similar to that of the black hole at the center of the nearby giant elliptical galaxy M87 (Harms et al. 1994, Macchetto et al. 1997). If the quasar were radiating below the Eddington limit, the inferred black hole mass would be even higher. In the Elvis et al. (1994) template, ≈ 1% of the bolometric luminosity is emitted in the observed z ′ band. If this fraction is larger for SDSS 1044-0125, the implied black hole mass would be reduced. Note that the universe was less than 1 Gyr old at this redshift, while the Eddington time scale, the e-folding time for the growing of a black hole shining at the Eddington luminosity, is 4 × 107(ǫ/0.1) yr, where ǫ is the radiative efficiency for the accretion. If the black hole started accreting with an initial mass of ∼ 103M⊙ with 10% efficiency, the seed black hole would have to form and begin accreting at redshift well beyond 10 in order to grow to 3 × 109M⊙ at z = 5.80 (see also Turner 1991). Forming such a massive black hole in such a short time is a remarkable feat. The observations of high-redshift quasars can be used to constrain the formation epoch of the first star clusters and the fueling process of early black holes. How likely is it to find a quasar like SDSS 1044-0125 in popular cold dark matter cosmological models? We use the following simple model to estimate the abundance of high-redshift quasars (see also Haiman & Hui 2000): Magorrian et al. (1998) have found a correlation between central black hole mass and bulge mass, MBH /Mbulge = 6 × 10−3 in nearby galaxies. If this correlation holds at high redshift, this would imply a bulge mass of – 12 – 5.7 × 1011 M⊙ (Λ-model) or 4.5 × 1011 M⊙ (Ω = 1) for the host galaxy of SDSS 1044-0125, and a lower limit of 5.7 × 1012 M⊙ (Λ-model) or 4.5 × 1012 M⊙ (Ω = 1) for its dark halo, assuming Mhalo/Mbulge ≥ ΩDM/Ωb ≈ 10. The comoving abundance of dark matter halos at this epoch is very sensitive to this halo mass, and can be estimated by means of the Press & Schechter (1974) formalism. In a Λ-CDM model, assuming σ8 = 0.87 and an untilted primordial power spectrum, for parent halo masses of 1012 M⊙ , 6 × 1012 M⊙ , and 1013 M⊙ , we expect 50,000, 60, and 4 candidate halos respectively within the survey volume in a redshift window ∆z = 1. The duty cycle of quasar activity is poorly known, but is certainly much less than unity (Haiman & Hui 2000). Given these uncertainties, this model is not contradictory to our discovery of a single quasar like SDSS 1044-0125 in the 600 deg2 survey area. Note that this model does not address the physical process by which the massive black hole formed. The assumptions we made about MBH/Mbulge and the lifetime of the quasars are completely untested at high redshift. Indeed, Rix et al. (1999) argue that the host galaxies of z ∼ 2 quasars are appreciably less luminous than the universal MBH/Mbulge hypothesis would imply. Because of its high luminosity, SDSS 1044-0125 likely probes the exponential, high–mass tail of the underlying dark halo distribution, making predictions of the expected number counts also sensitive to cosmological parameters, especially the normalization of the power spectrum (i.e. σ8 ). In principle, the SDSS survey will be able to probe quasars ≈ 1 mag fainter than SDSS 1044-0125, and could reveal tens of additional sources at z ≈ 6. The detection of these ob jects will yield strong constraints on cosmological models for the formation and evolution of quasars at very high redshifts. The Sloan Digital Sky Survey (SDSS) is a joint pro ject of the University of Chicago, Fermilab, the Institute for Advanced Study, the Japan Participation Group, The Johns Hopkins University, the Max-Planck-Institute for Astronomy, Princeton University, the United States Naval Observatory, and the University of Washington. Apache Point Observatory, site of the SDSS, is operated by the Astrophysical Research Consortium. Funding for the pro ject has been provided by the Alfred P. Sloan Foundation, the SDSS member institutions, the National Aeronautics and Space Administration, the National Science Foundation, the U.S. Department of Energy, and Monbusho, Japan. The SDSS Web site is http://www.sdss.org/. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint pro ject of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by NASA and NSF. XF and MAS acknowledge additional support from Research Corporation, NSF grant AST96-16901, the Princeton University Research Board, and a Porter O. Jacobus Fellowship. RHB acknowledges support from the Institute of Geophysics and – 13 – Planetary Physics (operated under the auspices of the U.S. Department of Energy by the University of California Lawrence Livermore National Laboratory under contract No. W- 7405-Eng-48). ZH acknowledges support from Hubble Fellowship grant HF-01119.01-99A. DPS acknowledges support from NSF grant AST99-00703. We thank Wolfgang Voges, Hans-Walter Rix, David Weinberg, and Peng Oh for helpful comments, and the expert assistance of Bob Goodrich and Terry McDonald during the Keck observations. REFERENCES Becker, R. H., White, R. L., & Helfand, D. J. 1995, ApJ, 450, 559 Bi, H., Borner, G., & Chu, Y. 1992, A&A, 266, 1 Cen, R., Miralda-Escud´e, J., Ostriker, J. P., & Rauch, M. 1994, ApJ, 437, L9 Chen, H.-W., Lanzeta, K. M., & Pascarelle, S. 1999, Nature, 398, 586 Dey, A., Spinrad, H., Stern, D., Graham, J. R., & Chaffee, F. 1998, ApJ, 498, L93 Djorgovski, S. G., Odewahn, S. C., Gal, R. R., Brunner, R., de Carvalho, R. R. 1999, astro-ph/9908142 Elvis, M., Wilkes, B. J., McDowell, J. C., Green, R. F., Bechtold, J., Willner, S. P., Oey, M. S., Polomski, E., & Cutri, R. 1994, ApJS, 95, 1 Epps, H.W., & Miller, J.S. 1998, Proc. SPIE, 3355, 48 Fan, X. 1999, AJ, 117, 2528 Fan, X. et al. 1999, AJ, 118, 1 ——, 2000a, AJ, 119, 1 ——, 2000b, AJ, 119, 928 Francis, P.J., Hewett, P.C., Foltz, C.B., Chaffee, F.H., Weymann, R.J., & Morris, S.L. 1991, ApJ, 373, 465 Fukugita, M., Ichikawa, T., Gunn, J.E., Doi, M., Shimasaku, K., & Schneider, D.P. 1996, AJ, 111, 1748 Giallongo, E., D’Odorico, S., Fontana, A., McMahon, R. G., Savaglio, S., Cristiani, S., Molaro, P., & Trevese, D. 1994, ApJ, 425, L1 – 14 – Gunn, J. E., & Peterson, B. A. 1965, ApJ, 142, 1633 Gunn, J.E., et al. 1998, AJ, 116, 3040 Haehnelt, M. G., Natara jan, P., & Rees, M. J. 1998, MNRAS, 300, 817 Haiman, Z., & Hui, L. 2000, ApJ, submitted (astro-ph/0002190) Haiman, Z., & Loeb, A. 1998, ApJ, 503, 505 ——, 1999, ApJ, 519, 479 Hamann, F. & Ferland, G. 1999, ARA&A, 37, 487 Harms, R.J. et al. 1994, ApJ, 435, L35 Hernquist, L., Katz, N., Weinberg, D. H., & Miralda-Escud´e, J. 1996, ApJ, 457, L51 Hu, E. M., McMahon, R. G., & Cowie, L. L. 1999, ApJ, 522, L9 Kaspi, S., Brandt, W. N., & Schneider, D. P. 2000, AJ, in press (astro-ph/0001299) Kirkpatrick, J. D., et al. 1999, ApJ, 519, 802 Krauss, L., & Turner, M. 1995, Gen. Rel. Grav., 27,1137 Leggett, S. K., et al. 2000, ApJL, in press (astro-ph/0004408) Lupton, R.H., Gunn, J.E., & Szalay, A. 1999, AJ, 118, 1406 Macchetto, F., Marconi, A., Axon, D. J., Capetti, A., Sparks, W., & Crane, P. 1997, ApJ, 489, 579 Magorrian, J., et al. 1998, AJ, 115, 2285 Martini, P., & Weinberg, D. H. 2000, ApJ, submitted (astro-ph/0002384) Massey, P. 1988, ApJ, 328, 315 Massey, P., & Cronwall, C. 1990, ApJ, 358, 344 Matthews, K., & Soifer, B. T. 1994 in Infrared Astronomy with Arrays, ed. I. McLean, (Dorchecht: Kluwer), 239 Miralda-Escud´e, J. 1997, ApJ, 501, 15 Miralda-Escud´e, J., & Rees, M. J. 1993, MNRAS, 260, 617 – 15 – Oke J.B., & Korycansky D.G. 1982, ApJ, 255, 11 Ostriker, J. P., & Steinhardt, P. 1995, Nature, 377, 600 Persson, S. E., Murphy, D. C., Krzeminski, W., Roth, M., & Rieke, M. J. 1998, AJ, 116, 2475. Press, W. H., & Schechter, P. L. 1974, ApJ, 181, 425 Rees, M. 1999, in After the Dark Ages: When Galaxies were Young (the Universe at 2 < z < 5), ed. S. S. Holt and E. P. Smith, (AIP Press), 13 Rix, H.W., Falco, E., Impey, C., Kochanek, C., Lehar, J., McLeod, B., Munoz, J., & Peng, C. 1999, preprint (astro-ph/9910190) Schlegel, D.J, Finkbeiner, D.P., & Davis, M. 1998, ApJ, 500, 525 Schmidt, M., Schneider D.P., & Gunn, J.E. 1995, AJ, 110, 6 Schmidt, M., van Gorkom, J.H., Schneider, D.P., & Gunn, J.E. 1995, AJ, 109, 473 Schneider, D.P, Schmidt, M., & Gunn, J.E. 1991, AJ, 101, 2004 Schneider, D. P., et al. 2000a, PASP, 112, 6 Songaila, A., Hu, E. M., Cowie, L., L., & McMahon, R. G. 1999, ApJ, 525, L5 Spinrad, H., Stern, D., Bunker, A.J., Dey, A., Lanzetta, K., Yahil, A., Pascarelle, S., & Fern´andez-Soto, A. 1998, AJ, 116, 2617 Steidel, C.C., Adelberger, K.L., Dickinson, M., Giavalisco, M., Pettini, M, & Kellogg, M.A. 1998, ApJ, 492, 428 Steidel, C.C., Adelberger, K.L., Giavalisco, M., Dickinson, M., Pettini, M. 1999, ApJ, 519, 1 Stern, D., & Spinrad, H. 1999, PASP, 111, 1475 Stern, D., Spinrad, H., Eisenhardt, P., Bunker, A. J., Dawson, S., Stanford, S., & Elston, R. 2000, ApJ, 533, L75 Storrie-Lombardi, L.J., McMahon, R. G., Irwin, M.J., & Hazard, C. 1996, ApJ, 468, 121 Strauss, M. A., et al. 1999, ApJ, 522, L61 Tsvetanov, Z. I., et al. 2000, ApJ, 531, L61 – 16 – Turner, E. L. 1991, AJ, 101, 5 Voges, W., et al. 1999, A&A, 349, 389 van Breugel, W., De Breuck, C., Stanford, S. A., Stern, D., Rottgering, H., & Miley, C. 1999, ApJ, 518, L61 Weymann, R.J., Stern, D., Bunker, A., Spinrad, H., Chaffee, F.H., Thompson, R.I., & Storrie-Lombardi, L.J. 1998, ApJ, 505, L95 York, D. G., et al. 2000, AJ, submitted Zheng, W., et al. 2000, AJ, submitted (astro-ph/0005247) Zuo, L. 1993, A&A, 278, 343 This preprint was prepared with the AAS LATEX macros v4.0. – 17 – Table 1. SDSS and K-band Photometry of SDSSp J10:44:33.04 –01:25:02.2 u∗ 22.91 ± 0.53 g ∗ 23.91 ± 0.49 r∗ 25.13 ± 0.62 i∗ 21.81 ± 0.19 z ∗ 19.23 ± 0.07 K 17.02 ± 0.04 The SDSS photometry (u∗ , g ∗ , r∗ , i∗ , z ∗ ) is reported in terms of asinh magnitudes on the AB system. The asinh magnitude system is defined in Lupton, Gunn & Szalay (1999); it becomes a linear scale in flux when the absolute value of the signal-to-noise ratio is less than about 5. In this system, zero flux corresponds to 24.24, 24.91, 24.53, 23.89, and 22.47, in u∗ , g ∗ , r∗ , i∗ , and z ∗ , respectively; larger magnitudes refer to negative flux values. The K band photometry is on the Vega-based system. E (B − V ) in this direction is 0.054, from Schlegel et al. 1998 Table 2. Emission Line Properties Line λ (A) redshift EW (A, rest frame) OI+SiII 1302 SiIV+OIV] 1400 8858.5 ± 3.8 9507.3 ± 3.7 5.802 ± 0.003 5.789 ± 0.003 3.4 ± 0.5 7.0 ± 0.4 – 18 – 104433.04-01252.2 North i’ filter East z’ filter 60" Figure 1. Finding chart for SDSS 1044–0125 (discovery image from the SDSS). The field is 160′′ on a side. The field is given in both the i′ and z ′ bands (54.1s exposure time). – 19 – Figure 2. Optical spectrum of SDSS 1044–0125 observed with KeckII/ESI. The total exposure time is 3600s. The spectrum is smoothed to 4 A/pixel. The spectral resolution is ∼ 8 A at λ = 9000A. – 20 – Figure 3. ESI Echellete spectrum over the range 7300 – 8800 A. The spectrum has been smoothed to a resolution of 2 A. The dotted line shows a fν ∝ ν −0.5 continuum, normalized to the region 8650 – 8730 A. The dashed line indicates the region with minimum absorption (τ ∼ 0.4, see §3.1). The lower panel shows the noise level.
astro-ph/0610259
1
0610
2006-10-09T20:59:43
Magnetic activity on AB Doradus: Temporal evolution of starspots and differential rotation from 1988 to 1994
[ "astro-ph" ]
Surface brightness maps for the young K0 dwarf AB Doradus are reconstructed from archival data sets for epochs spanning 1988 to 1994. By using the signal-to-noise enhancement technique of Least-Squares Deconvolution, our results show a greatly increased resolution of spot features than obtained in previously published surface brightness reconstructions. These images show that for the exception of epoch 1988.96, the starspot distributions are dominated by a long-lived polar cap, and short-lived low to high latitude features. The fragmented polar cap at epoch 1988.96 could indicate a change in the nature of the dynamo in the star. For the first time we measure differential rotation for epochs with sufficient phase coverage (1992.05, 1993.89, 1994.87). These measurements show variations on a timescale of at least one year, with the strongest surface differential rotation ever measured for AB Dor occurring in 1994.86. In conjunction with previous investigations, our results represent the first long-term analysis of the temporal evolution of differential rotation on active stars.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- ?? (2006) Printed 28 June 2018 (MN LATEX style file v2.2) Magnetic activity on AB Doradus: Temporal evolution of starspots and differential rotation from 1988 to 1994 S.V. Jeffers1,2, J.-F. Donati 1, A.Collier Cameron2 1Laboratoire d'Astrophysique Toulouse-Tarbes, Observatoire Midi-Pyr´en´ees, 14, avenue Edouard Belin, F-31400 Toulouse, France 2School of Physics and Astronomy, University of St Andrews, North Haugh, St Andrews, Fife KY16 9SS, UK 6 0 0 2 t c O 9 1 v 9 5 2 0 1 6 0 / h p - o r t s a : v i X r a ABSTRACT Surface brightness maps for the young K0 dwarf AB Doradus are reconstructed from archival data sets for epochs spanning 1988 to 1994. By using the signal-to-noise enhancement tech- nique of Least-Squares Deconvolution, our results show a greatly increased resolution of spot features than obtained in previously published surface brightness reconstructions. These im- ages show that for the exception of epoch 1988.96, the starspot distributions are dominated by a long-lived polar cap, and short-lived low to high latitude features. The fragmented po- lar cap at epoch 1988.96 could indicate a change in the nature of the dynamo in the star. For the first time we measure differential rotation for epochs with sufficient phase coverage (1992.05, 1993.89, 1994.87). These measurements show variations on a timescale of at least one year, with the strongest surface differential rotation ever measured for AB Dor occurring in 1994.86. In conjunction with previous investigations, our results represent the first long- term analysis of the temporal evolution of differential rotation on active stars. Key words: stars: activity, spots, individual (AB Dor), rotation; line:profiles 1 INTRODUCTION Solar-type stars exhibit signatures of magnetic activity that are as- sumed to be based on dynamo mechanisms operating in the star's outer convective zone. The detailed workings of the generation and amplification mechanisms of the stellar dynamos are still poorly understood as a result of the complex physics involved in theoret- ical models, and the lack of observational constraints. However, it is widely accepted that differential rotation and convection are essential ingredients of common theoretical amplification models (Parker 1955; Babcock 1965; Leighton 1964, 1969). Differential rotation results from the interplay of rotation and convection, which leads to a redistribution of heat and angular momentum inside the convection zone. The thermal and density profiles of convective motions produce a dependence of differen- tial rotation on both stellar latitude and radius. It is a key pro- cess in the cyclic activity of the stellar dynamo being the process through which poloidal-toroidal field conversion occurs. To date there have been numerous measurements of differential rotation on rapidly rotating cool stars, summarised by Barnes et al. (2005), that show a steady increase in the magnitude of differential rota- tion towards earlier spectral types, consistent with the theoretical predictions of Rudiger & Kuker (2002). In addition, helioseismol- ogy reveals a differential rotation profile that varies with radius and latitude and is caused by the presence of an intense turbulence in the solar convention zone, which is driven by Reynolds stresses (Brun & Toomre 2002). c(cid:13) 2006 RAS Several methods have been used to measure differential rotation. The methods of Gray (1977); Bruning (1981) and Reiners & Schmitt (2002) use Fourier analysis to detect differen- tial rotation through line profile analysis. These methods are only applicable to stars that do not exhibit large cool starspots as they distort the shape of the line profile. Other methods reconstruct sur- face brightness images over a time period and then trace surface features to ascertain how their rotation periods are dependant on latitude. Examples include, the cross-correlation method used by Donati & Collier Cameron (1997) on AB Dor to obtain the first dif- ferential rotation measurement for a star other than the Sun. This method measures the amount of rotational shear as a function of latitude by cross-correlating belts of equal latitude. The sheared- image method extends this to include a solar-like differential rota- tion law into the image reconstruction process, where the rotation rate is allowed to vary smoothly with latitude on an image grid (Donati et al. 2000). The temporal evolution of differential rotation can be deter- mined by measurements over several epochs as shown for AB Dor by Collier Cameron & Donati (2002). In this work, a matched-filter analysis method was used to track individual spot features in the trailed spectrograms. The temporal evolution of differential rota- tion on AB Dor has also been confirmed by Donati et al. (2003), through the use of the sheared-image method, though not for the same epochs as Collier Cameron & Donati (2002). In this paper we extend the epochs for which differential ro- tation has been measured using the sheared-image method, by pro- 2 S.V. Jeffers, J.-F. Donati, A.Collier Cameron Epoch Spectral Domain (nm) Pixel Size (km s−1) Resolution Telescope/Spectrograph 1988 Dec 16,19 1988 Dec 21 1992 Jan 18,19,20 1992 Dec 14 1993 Nov 23,24,25 1994 Nov 15,16,17 605.9249 - 724.4182 581.0809 - 673.5818 498.2082 - 708.8381 452.9687 - 688.6151 479.6900 - 751.7410 551.0686 - 792.9818 2.7 4.6 2.42 2.98 2.98 3.56 56000 27000 51100 33600 33600 26600 AAT/UCL ESO(3.6m)/CASPEC AAT/UCL AAT/UCL AAT/UCL CTIO Table 1. Journal of observations for AB Dor showing the observed spectral domain, pixel size, resolution and the instrument used for each epoch. Epoch 1988.96 1988.97 1992.05 1992.95 1993.89 1994.87 Julian Date Phase Range Signal-to-noise No of lines References 7511.7607 - 7514.8768 7516.5617 - 7518.8339 8639.9234 - 8642.2515 8970.9158 - 8971.2737 9314.9249 - 9317.2631 9671.5421 - 9673.8443 -0.37-(-0.174) : 0.381-0.672 0.961-1.517 0.127-0.782 : 0.038-0.722 : 0.98-0.628 0.092-0.766 -0.656 -- 0.034 : -0.266-0.918 : 0.175-0.886 0.087-0.591 : 0.698-0.525 : 0.604-0.555 70-100 280-300 50-110 110-180 56-95 160-250 80 290 945 2134 1515 990 Collier Cameron et al. (1990) Collier Cameron et al. (1990) Collier Cameron & Unruh (1994) Collier Cameron (1995) Unruh et al. (1995) Collier Cameron et al. (1999) Table 2. Journal of observations showing Julian dates (+2450000), phase coverage, signal-to-noise of the data set, and the number of lines used in deconvolu- tion. cessing archival AB Dor data for epochs December 1988, Jan- uary 1992, November 1993 and November 1994 (presented in Sec- tion 2). This is the first time that surface brightness images have been reconstructed from composite profiles computed from these data sets using the signal-to-noise enhancement technique Least- Squares Deconvolution (LSD) (presented in Sections 3 and 4). Fi- nally we measure differential rotation for each eopch with sufficient rotational phase coverage (presented in Section 5) and discuss the implications of our results in Section 6. 2 OBSERVATIONS AND DATA MODELLING The details of the instrument configuration and observing proce- dures used to secure the six data sets are summarised in Table 1. We refer the reader to the publications listed in Table 2 for further details. 2.1 Data Reduction All frames were processed with ESpRIT, a dedicated package for the optimal extraction of echelle spectroscopic observations. Firstly a 3-D fit of the bias frame is subtracted from the flat-field and arc frames. Each order of the raw echelle frame is then located and traced using cross-correlations with a user defined reference profile. A linear or 2D fit to the shape of the arc lines provides the slit direction. Deviations from the slit direction are averaged over all orders provide the mean slit shape. The first step in the wavelength calibration procedure is to obtain an accurate identification of calibration lines. To start, the user specifies the order numbers, and approximate values for the wavelength of the first pixel and pixel size. This information is used to determine the location of lines from a calibration line list. A quadratic dispersion polynomial is then determined and used to calibrate the remaining orders. The final dispersion relation is ob- tained by fitting a 2D polynomial to the pixel positions and corre- sponding wavelengths of all lines simultaneously. The comprising dimensions of the polynomial fit are; 1 dimension to fit the disper- sion relation of each order and 1 dimension to fit its variation from one order to the next. Pixel-to-pixel sensitivity differences are removed by dividing each pixel in the stellar frame by the corresponding pixel in the flat-field. A 2D polynomial fit to the inter order background is sub- tracted from the stellar frame. All pixels that deviate from the av- erage intensity of the order (i.e. cosmic rays) are removed. The op- timal extraction of Marsh (1989) is then implemented. The contin- uum is automatically fitted firstly by using a high-degree 1D poly- nomial, and then by a 2D polynomial to identify any systematic trends in the continuum's shape. A more detailed description of this package is given by Donati et al. (1997). 2.2 Least-Squares Deconvolution LSD is a method for combining the rotation profiles of thousands of spectral lines in an optimally weighted manner (Donati et al. 1997). It uses a weighted least squares algorithm to compute the line broadening profile which, when convolved with the known pattern of photospheric absorption lines in a stellar spectrum, op- timises the chi-squared fit to the data. The list of spectral lines is obtained from the LTE model atmospheres of Kurucz (1993) for Tef f =5000 K and log g = 4.5, where features with a relative central depth of at least 40% of the local continuum flux are used. The total number of lines used for each epoch is shown in Table 2. It can be shown that the enhancement in signal to noise is equivalent to either an optimally-weighted stacking of the profiles, or to cross-correlation with the line pattern, scaling as the square root of the number of lines used. It has the advantage, however, that sidelobes caused by blends are automatically eliminated from the composite profile, giving a clean profile surrounded by flat con- tinuum. LSD conserves the shape of the rotational profile, implying that any deviations in this profile can be interpreted as brightness inhomogeneities on the stellar surface. c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 3 1.0 c I / I 0.5 -0.200 -0.198 -0.195 -0.193 -0.191 -0.189 -0.187 -0.179 -0.176 -0.174 -0.172 -0.169 -0.167 -0.165 -0.163 -0.160 -0.158 -0.148 -0.145 -0.143 -0.141 -0.139 -0.136 -0.134 -0.132 -0.100 -0.098 -0.096 -0.094 -0.092 -0.089 -0.087 -0.085 -0.083 -0.080 -0.078 -0.076 -0.074 -0.072 -0.069 -0.064 -0.053 -0.042 -0.040 -0.035 -0.033 -0.025 -0.021 -0.018 -0.006 -0.004 0.551 0.553 0.555 0.557 0.559 0.568 0.571 0.591 0.593 0.595 0.598 0.606 0.608 0.621 0.623 0.637 0.639 0.648 0.650 0.652 0.660 0.662 0.687 0.689 0.691 0.702 0.706 0.717 0.719 0.721 0.742 0.744 0.746 0.767 0.769 0.771 0.782 0.787 0.796 0.798 0.800 0.810 0.814 0.816 0.825 0.827 0.829 0.840 0.842 0.0 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 2. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 16+19 December 1988, observed at the AAT. The rotational phases are indicated to the right of each profile. 3 RADIAL VELOCITY CORRECTION 3.1 Telluric line alignment Telluric lines are used to correct for small shifts in the spectro- graph during the night i.e. from dewar refill and the thermal and c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? mechanical relaxation of spectrograph's components, as they are only present in the Earth's atmosphere and are therefore at zero ra- dial velocity. We use the procedure of Donati et al. (2003) in this analysis. Firstly the composite profile of telluric absorption lines in each stellar spectrum is computed using LSD, with a line mask comprising the wavelengths and relative strengths of known telluric 4 S.V. Jeffers, J.-F. Donati, A.Collier Cameron 1.0 0.961 0.969 0.086 0.094 0.205 0.212 0.315 0.323 0.977 0.102 0.221 0.331 0.985 0.992 0.110 0.125 0.229 0.236 0.338 0.347 0.000 0.133 0.244 0.354 0.008 0.016 0.141 0.149 0.252 0.260 0.362 0.370 c I / I 0.5 0.031 0.158 0.268 0.379 0.047 0.054 0.165 0.173 0.276 0.284 0.509 0.517 0.062 0.181 0.291 0.070 0.189 0.299 0.078 0.197 0.307 0.0 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 3. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 21 December 1988, observed at ESO. The rotational phases are indicated to the right of each profile. features (mainly of water) in the observed spectral region. The in- strumental velocity of the composite telluric profile is then mea- sured and used to define the zero-point of the velocity scale. Each stellar profile is then shifted by the measured instrumental velocity before applying heliocentric velocity corrections. 3.2 Radial Velocity Results The radial velocity derived for each epoch is tabulated in Table 1, where over the time-span of our observations there is a variation of 3.4 km s−1. The change in radial velocity reflects the orbital motion of AB Dor predominantly due to the presence of its closest com- panion, AB Dor C (Close et al. (2005)). AB Dor C is a low mass object (0.09 M⊙) first detected by Guirado J.C. et al. (1997), which orbits AB Dor A at a separation of 0.156 arcseconds. In addition to AB Dor C, AB Dor is known to have a wide companion AB Dor B (Vilhu et al. 1991; Martin & Davey 1995) which is itself a close binary system, and is separated from AB Dor A by a distance of 9.09±0.01 arcseconds. The AB Dor A and C orbital solution has been deter- mined by Close et al. (2005), where the parameters that define AB Dor A's reflex orbit are fitted by: period = 11.75 ± 0.25 yr, semi-major axis = 0.032 ±0.002′′, eccentricity = 0.59 ±0.03, peri- astron passage at 1991.8 ± 0.2, inclination = 67 ±3◦, Ω=132 ±2◦, ω=107 ±7◦. The orbital solution is shown in Fig. 1 to fit the data points within the error bars and to follow the decrease in radial velocity at periastron (epoch 1991.8). The error is approximately 0.5 km s−1 for the epochs of this work and 0.3 km s−1 for epochs after 1995, and results from broad and time variable stellar lines. The exceptions are the radial velocity measurements of epochs 1992.95, 1993.89, and 1994.87, but the good fits to the line pro- file shown in Figures 5 & 6 for epoch 1992.05, Figure 9 for epoch 1992.95 and Figures 10 & 11 for epoch 1993.89, exclude any addi- tional measurement errors. More empirical data points will help to constrain the errors of the orbital solution. 4 SURFACE IMAGE RECONSTRUCTION The surface brightness images are reconstructed using the max- imum entropy code of Brown et al. (1991) and Donati & Brown (1997). The brightness model that is incorporated in this code is c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 5 1 0.8 0.6 0.4 0.2 0 All Latitudes Latitudes 0o to 50o 90 80 70 60 50 40 30 20 10 0.002 0.001 Frac Spot Coverage 0 0 e g a r e v o C t o p S . c a r F 0.18 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 Figure 4. Maximum Entropy surface brightness distribution for December 1988 (epoch 1988.96), where the vertical tics at the top of the plot indicate the phase coverage. The plot to the left shows the fractional spot coverage per latitude bin integrated over longitude, while the plot below shows the fractional spot coverage per rotational phase bin, integrated over latitude. 33 32.5 32 31.5 31 30.5 30 29.5 29 s / m k y t i c o e V l l i a d a R Epoch Radial Velocity (km s−1) 1988 Dec 16 & 19 1992 Jan 1992 Dec 1993 Nov 1994 Nov 1995 Dec 1996 Dec 1998 Jan 1998 Dec 1999 Dec 2000 Dec 2001 Dec 2002 Dec 32.2 31.2 28.8 29.7 29.9 31.4 31.4 31.5 31.6 31.8 32.1 32.7 32.9 28.5 1988 1990 1992 1994 1996 1998 2000 2002 2004 Date of Observation Figure 1. The orbital solution of Close et al. (2005) for AB Dor A's reflex orbit. Also plotted are the empirically derived radial velocity values for AB Dor as determined by this analysis (epochs 1988 to 1994), The orbital solu- tion includes an offset in the y-axis of 31.37 km s−1 that corresponds to the radial velocity of the binary system, which was determined by χ2 minimi- sation. The χ2 obtained for this solution is 6.023. the 'spot occupancy' model of Collier Cameron (1992), where each point on the stellar surface is quantified by the local fraction of the stellar surface occupied by spots. The range of spot occupancy is from 0, where there are no spots present, to 1, where there is maxi- mum spottedness. The imaging parameters that are used in this work for AB Dor c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Table 3. Radial velocity measurements for each epoch of this analysis with previously published values for epochs after 1994. are stellar axial inclination, i = 60o, the projected equatorial rotation velocity, v sin i = 89 km s−1 and the photospheric and spot tempera- tures respectively 5000 K and 3500 K (Donati et al. 2003). All data sets are phased according to the ephemeris of Innis et al. (1988); HJD = 244 4296.575 + 0.51479 E. LSD profiles of slowly rotating standard stars (GL 176.3 and GL 367) are used as template profiles that represent the contribution of the photosphere and the spot to the shape of the intrinsic line profile. 6 S.V. Jeffers, J.-F. Donati, A.Collier Cameron 1.0 c I / I 0.5 0.0 0.127 0.131 0.135 0.140 0.144 0.148 0.153 0.157 0.161 0.166 0.170 0.175 0.179 0.183 0.188 0.192 0.196 0.201 0.205 0.209 0.214 0.218 0.222 0.227 0.231 0.236 0.238 0.244 0.249 0.253 0.257 0.262 0.266 0.270 0.275 0.279 0.281 0.284 0.292 0.297 0.299 0.305 0.307 0.314 0.319 0.323 0.325 0.332 0.336 0.338 0.356 0.361 0.365 0.369 0.374 0.378 0.382 0.387 0.391 0.395 0.398 0.404 0.409 0.411 0.413 0.422 0.424 0.431 0.435 0.439 0.444 0.448 0.450 0.457 0.459 0.466 0.470 0.474 0.479 0.483 0.485 0.492 0.496 0.500 0.505 0.509 0.511 0.518 0.522 0.527 0.531 0.535 0.540 0.542 0.549 0.553 0.557 0.559 0.566 0.570 0.575 0.579 0.583 0.588 0.592 0.597 0.601 0.605 0.610 0.614 0.616 0.623 0.625 0.627 0.636 0.640 0.644 0.649 0.651 0.653 0.662 0.664 0.671 0.673 0.680 0.684 0.688 0.693 0.695 0.701 0.706 0.710 0.712 0.719 0.723 0.728 0.732 0.736 0.741 0.745 0.747 0.754 0.758 0.763 0.767 0.771 0.776 0.780 0.782 0.038 0.043 0.047 0.052 0.054 0.060 0.065 0.069 0.073 0.078 0.082 0.086 0.091 0.131 0.152 0.205 0.251 0.255 0.260 0.264 0.268 0.273 0.277 0.281 0.286 0.290 0.294 0.299 0.303 0.307 0.312 0.316 0.320 0.325 0.329 0.334 0.338 0.342 0.365 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 5. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 18-20 January 1992, part I. The rotational phases are indicated to the right of each profile. 4.1 Results All data sets provide good sampling of the rotational cycle of AB Dor. The maximum entropy images are structurally very similar with a polar cap and/or high latitude spots, and with varying de- grees of low latitude spots. The longitude resolution is approxi- mately 3◦ at the equator and the size of the smallest features in latitude. The first surface brightness image reconstructed for December 1988 is shown in Figure 4. It shows high latitude spots that domi- nate over a weak polar cap and a dearth of low latitude spot cover- age. These sets were originally observed with the aim of detecting c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 7 1.0 c I / I 0.5 0.0 0.370 0.374 0.379 0.381 0.387 0.400 0.405 0.407 0.413 0.418 0.422 0.427 0.431 0.435 0.440 0.444 0.448 0.453 0.457 0.461 0.466 0.470 0.474 0.477 0.483 0.488 0.492 0.496 0.501 0.505 0.507 0.514 0.518 0.523 0.527 0.570 0.574 0.579 0.583 0.587 0.592 0.596 0.598 0.605 0.609 0.613 0.618 0.622 0.627 0.631 0.633 0.640 0.644 0.646 0.653 0.657 0.661 0.666 0.670 0.675 0.679 0.683 0.688 0.692 0.696 0.701 0.705 0.709 0.714 0.718 0.722 0.980 0.984 0.989 0.993 0.997 0.002 0.006 0.010 0.015 0.017 0.024 0.028 0.030 0.037 0.041 0.045 0.050 0.054 0.058 0.063 0.067 0.072 0.076 0.080 0.085 0.089 0.093 0.098 0.102 0.106 0.111 0.115 0.144 0.148 0.152 0.157 0.159 0.165 0.170 0.174 0.178 0.183 0.196 0.200 0.205 0.222 0.226 0.230 0.235 0.239 0.243 0.248 0.252 0.256 0.261 0.265 0.269 0.274 0.278 0.283 0.287 0.291 0.296 0.300 0.304 0.309 0.313 0.317 0.322 0.326 0.330 0.335 0.339 0.357 0.362 0.366 0.370 0.375 0.379 0.383 0.388 0.392 0.396 0.401 0.405 0.410 0.414 0.418 0.423 0.427 0.431 0.436 0.440 0.444 0.449 0.453 0.457 0.462 0.466 0.470 0.475 0.479 0.484 0.488 0.492 0.497 0.528 0.530 0.537 0.541 0.554 0.628 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 6. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 18-20 January 1992, part II. The rotational phases are indicated to the right of each profile. circumstellar clouds on AB Dor using Hα, Ca II H and K, and Mg II h and k lines Collier Cameron et al. (1990). The most striking feature of this data set is the fragmented polar cap. However, this could result from the low S/N of the data set. The reconstructed surface-brightness image for January 1992 is shown in Figure 7. The image is comparable with images of previous this data set processed without LSD Collier Cameron & Unruh (1994) using the Ca I 643.9nm and Ca 671.8 nm photospheric lines. It is not possible to compare exact fea- tures due to the signal-to-noise difference, but to a first approxima- tion the two images contain similar spot features at mid to low lat- itudes. Examples of such features are at phase 0.65, the gap in low c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? 8 S.V. Jeffers, J.-F. Donati, A.Collier Cameron 90 80 70 60 50 40 30 20 10 0 0.0025 0.002 0.0015 0.001 0.0005 0 Frac Spot Coverage e g a r e v o C t o p S . c a r F 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 1 0.8 0.6 0.4 0.2 0 All Latitudes Latitudes 0o to 50o Figure 7. Maximum Entropy surface brightness distribution for January 1992 (epoch 1992.05), where the vertical tics at the top of the plot indicate the phase coverage. The plot to the left shows the fractional spot coverage per latitude bin integrated over longitude, while the plot below shows the fractional spot coverage per rotational phase bin, integrated over latitude. 90 80 70 60 50 40 30 20 10 0 0 0.002 0.001 Frac Spot Coverage e g a r e v o C t o p S . c a r F 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 1 0.8 0.6 0.4 0.2 0 All Latitudes Latitudes 0o to 50o Figure 8. Maximum Entropy surface brightness distribution for December 1992 (epoch 1992.95), where the vertical tics at the top of the plot indicate the phase coverage. The plot to the left shows the fractional spot coverage per latitude bin integrated over longitude, while the plot below shows the fractional spot coverage per rotational phase bin, integrated over latitude. c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 9 1.0 c I / I 0.5 0.092 0.095 0.101 0.103 0.110 0.114 0.123 0.125 0.131 0.136 0.140 0.144 0.168 0.172 0.176 0.181 0.183 0.189 0.191 0.198 0.202 0.206 0.210 0.213 0.220 0.224 0.226 0.232 0.234 0.241 0.271 0.275 0.279 0.283 0.287 0.290 0.296 0.300 0.305 0.309 0.313 0.317 0.321 0.324 0.330 0.332 0.339 0.343 0.345 0.351 0.356 0.358 0.364 0.368 0.406 0.410 0.414 0.420 0.424 0.429 0.433 0.437 0.441 0.446 0.450 0.452 0.466 0.469 0.475 0.479 0.481 0.488 0.492 0.494 0.501 0.505 0.509 0.511 0.518 0.522 0.524 0.531 0.535 0.539 0.544 0.548 0.552 0.556 0.561 0.565 0.585 0.589 0.593 0.597 0.601 0.604 0.610 0.614 0.616 0.623 0.627 0.629 0.649 0.653 0.658 0.662 0.666 0.671 0.673 0.679 0.683 0.688 0.692 0.696 0.700 0.702 0.709 0.713 0.717 0.722 0.726 0.728 0.734 0.739 0.741 0.747 0.766 0.0 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 9. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 14 December 1992. The rotational phases are indicated to the right of each profile. to mid-latitude features at phase 0.35, and the spot features at phase 0.5. Collier Cameron & Unruh (1994) also reconstructed a surface brightness image using the photospheric line Fe I 666.3nm. How- ever, the reconstructed spot features and groupings are less simi- lar to the image shown in Figure 7, which is likely to result from the different excitations of the two lines. At high to polar latitudes the images reconstructed in this analysis show a strong and uni- form polar cap whereas the images of Collier Cameron & Unruh (1994) show weak and fragmented spot structures. It should be noted that the default value of spot coverage is 0.5 in the work of Collier Cameron & Unruh (1994), while in this analysis we set the value to be 0.999. A surface image has also been reconstructed c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? 10 S.V. Jeffers, J.-F. Donati, A.Collier Cameron 1.0 c I / I 0.5 0.0 -0.656 -0.652 -0.641 -0.639 -0.637 -0.626 -0.622 -0.620 -0.609 -0.579 -0.575 -0.571 -0.569 -0.567 -0.565 -0.553 -0.549 -0.547 -0.541 -0.538 -0.536 -0.509 -0.505 -0.503 -0.497 -0.494 -0.488 -0.484 -0.479 -0.475 -0.471 -0.467 -0.464 -0.440 -0.436 -0.432 -0.430 -0.428 -0.425 -0.423 -0.421 -0.419 -0.417 -0.398 -0.396 -0.368 -0.366 -0.359 -0.346 -0.344 -0.338 -0.336 -0.334 -0.331 -0.298 -0.296 -0.294 -0.285 -0.283 -0.281 -0.279 -0.276 -0.274 -0.259 -0.257 -0.255 -0.227 -0.225 -0.222 -0.220 -0.218 -0.216 -0.214 -0.212 -0.193 -0.188 -0.184 -0.180 -0.158 -0.156 -0.154 -0.152 -0.150 -0.137 -0.132 -0.128 -0.126 -0.124 -0.115 -0.113 -0.093 -0.089 -0.087 -0.085 -0.083 -0.072 -0.070 -0.068 -0.066 -0.055 -0.051 -0.049 -0.042 -0.038 -0.036 -0.034 0.266 0.271 0.273 0.279 0.281 0.283 0.286 0.296 0.298 0.301 0.303 0.305 0.336 0.340 0.342 0.353 0.355 0.361 0.363 0.366 0.368 0.423 0.425 0.427 0.433 0.436 0.440 0.442 0.444 0.471 0.475 0.477 0.484 0.488 0.492 0.495 0.497 0.499 0.501 0.514 0.516 0.541 0.543 0.545 0.554 0.558 0.562 0.565 0.567 0.569 0.581 0.583 0.585 0.619 0.623 0.625 0.632 0.634 0.636 0.638 0.640 0.642 0.644 0.661 0.666 0.686 0.691 0.695 0.697 0.699 0.708 0.710 0.712 0.714 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 10. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 23-25 November 1993, part I. The rotational phases are indicated to the right of each profile. by Jarvinen et al. (2005) using photometric data for the epoch 1991.96, which shows a primary spot at phase 0.7 and a weaker secondary spot at phase 1.097. While Figure 7 shows that there is a fragmented spot structure at phases 0.6 to 0.8, there is no evidence for a second grouping of spots at phase 1.097. For the single night of observations in December 1992 the re- constructed surface brightness image is broadly in agreement with that previously reconstructed by Collier Cameron (1995) using Ca I 643.9 nm, Fe I 666.3 nm and Ca I 671.8 nm. Both images show an off-centre polar cap though the spot structure is less fragmented, and the lower latitude features are more clearly resolved in the sur- face brightness images reconstructed in this analysis. Examples of c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 11 1.0 c I / I 0.5 0.0 0.716 0.719 0.721 0.754 0.756 0.762 0.764 0.771 0.775 0.777 0.779 0.781 0.783 0.786 0.801 0.822 0.827 0.829 0.831 0.833 0.835 0.837 0.840 0.842 0.844 0.865 0.867 0.886 0.890 0.895 0.897 0.899 0.901 0.903 0.905 0.907 0.909 0.912 0.914 0.916 0.918 0.175 0.179 0.181 0.183 0.186 0.196 0.201 0.203 0.205 0.207 0.209 0.211 0.213 0.215 0.218 0.220 0.222 0.224 0.226 0.272 0.276 0.278 0.285 0.287 0.293 0.296 0.298 0.300 0.302 0.304 0.306 0.340 0.345 0.347 0.349 0.357 0.360 0.366 0.368 0.370 0.379 0.383 0.387 0.407 0.409 0.411 0.420 0.424 0.426 0.428 0.431 0.433 0.435 0.437 0.439 0.475 0.477 0.479 0.482 0.492 0.497 0.499 0.501 0.503 0.505 0.507 0.509 0.542 0.546 0.548 0.550 0.559 0.563 0.565 0.567 0.576 0.578 0.584 0.589 0.611 0.615 0.619 0.621 0.623 0.626 0.628 0.630 0.632 0.634 0.653 0.658 0.676 0.680 0.684 0.686 0.688 0.690 0.693 0.705 0.708 0.710 0.718 0.722 0.743 0.748 0.752 0.756 0.758 0.760 0.762 0.765 0.767 0.769 0.771 0.790 0.807 0.811 0.813 0.815 0.824 0.826 0.828 0.830 0.832 0.845 0.847 0.850 0.852 0.854 0.856 0.858 0.860 0.862 0.884 0.886 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 11. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 23-25 November 1993, part II. The rotational phases are indicated to the right of each profile. common features include spots at phases 0.15, 0.28 and 0.45, and a large unspotted area at phase 0.35. For epoch 1992.96 the photo- metric image reconstruction of Jarvinen et al. (2005) shows a pri- mary spot at phase 0.931, and a secondary spot at phase 1.325. Due to missing phase coverage it is not possible to verify the position- ing of the primary spot from our Doppler images, but we show that there are no significant spot features present at phase 1.325. The surface brightness image for November 1993 is shown in Figure 12. These images were previously reconstructed using the the combined lines of Ca I 643.9 nm, Fe I 666.3 nm and Ca I 671.8 nm by Unruh et al. (1995). The images reconstructed in this c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? 12 S.V. Jeffers, J.-F. Donati, A.Collier Cameron 90 80 70 60 50 40 30 20 10 0 0.0025 0.002 0.0015 0.001 0.0005 0 Frac Spot Coverage e g a r e v o C t o p S . c a r F 0.25 0.2 0.15 0.1 0.05 0 1 0.8 0.6 0.4 0.2 0 All Latitudes Latitudes 0o to 50o Figure 12. Maximum Entropy surface brightness distribution for November 1993 (epoch 1993.89), where the vertical tics at the top of the plot indicate the phase coverage. The plot to the left shows the fractional spot coverage per latitude bin integrated over longitude, while the plot below shows the fractional spot coverage per rotational phase bin, integrated over latitude. work have more intermediate and low latitude spot features. How- ever, there are comparable spot features at phases 0.38 to 0.55, 0.62, 0.8 and a similar spot grouping at phase 0, though as indicated by the tick marks above the plot, there is no phase coverage in this re- gion. Similar to the comparisons of the 1992 data sets, at high and polar latitudes the spot structure is weaker and more fragmented in the images reconstructed by Unruh et al. (1995). The surface image of Jarvinen et al. (2005) at epoch 1993.89 shows a primary spot at phase 1.542 and a secondary spot at phase 1.097. At phase 1.097 in Figure 12 there is a large grouping of spots, while at phase 1.542 there is a weak spot feature that is insignificant in strength com- pared to other reconstructed spots at phases 0.38, 0.49 and 0.62. The reconstructed surface-brightness image for November 1994 is shown in Figure 15. LSD has previously been applied to the 1994 CTIO data by Collier Cameron et al. (1999). However, in our image reconstructions the polar cap is stronger and less fragmented and at mid to low latitudes the spot features are more resolved. Ex- amples of similar spot features include those at phase 0.1 to 0.18, 0.25 to 0.35 and 0.45 to 0.55. For epoch 1994.87 the surface image of Jarvinen et al. (2005) shows a primary spot at phase 1.514, and a secondary spot at phase 1.097. While there is a spot feature at phase 1.514, it is significantly weaker than its neighbouring features, and the spot feature at phase 0.3 in Figure 15. We have reconstructed a weak spot feature at the phase 1.097, but there are other stronger spots close by at phase 0.3. At certain phases of the LSD profiles, it is possible to distin- guish a shallow absorption feature migrating through the line pro- file, which is not reproduced in the model Maximum Entropy pro- files. Examples of such features are shown in part II of the Novem- ber 1993 data set (Figure 11) at phases 0.565 to 0.815. We attribute these absorption features as being small regions on the stellar sur- face that are brighter than the surrounding photosphere in contrast to cooler regions that produce emission features in the line profiles. As the imaging code is designed only to reconstruct cool features on the stellar surface, it is not possible to reconstruct these bright features. The presence of these bright features does not interfere with the reliable reconstruction of cool spots. 5 SURFACE DIFFERENTIAL ROTATION We used the sheared-image method of Donati et al (2000) to mea- sure the differential rotation of AB Dor. AB Dor is an ideal can- didate to measure differential rotation as its short rotation period (0.51479-d or 12.2053 rad d−1) means that it is possible to observe up to two-thirds of the stellar surface in one night and to get the necessary overlapping phase coverage within a few days. The image reconstruction process also incorporates a model of the stellar surface whose rotation rate Ω depends on latitude ac- cording to the simplified solar-like differential rotation law: Ω(θ) = Ωeq − δΩcos2θ (1) where Ω(θ) is the rotation rate at colatitude θ, Ωeq is the equa- torial rotation rate and δΩ is the difference between polar and equa- torial rotation rates. We performed a large set of of image recon- structions, using a two-dimensional grid of values for the param- eters Ωeq,δΩ. For each set of model parameters, the image recon- struction was driven until it reached a fixed value of spot filling factor. The χ2 values of the resulting images form a "landscape" on this grid. The best fitting model will corresponds to the mini- mum in the χ2 landscape, as models with the wrong shear will give poor fits to the data (Petit et al. 2002). The optimally fitting differential rotation parameters and their errors are determined by fitting the reduced χ2 landscape grids with c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 13 1.0 c I / I 0.5 0.087 0.095 0.099 0.104 0.108 0.112 0.116 0.120 0.124 0.132 0.136 0.153 0.157 0.162 0.166 0.170 0.174 0.178 0.182 0.186 0.190 0.198 0.213 0.221 0.226 0.230 0.234 0.238 0.242 0.246 0.250 0.254 0.258 0.263 0.267 0.271 0.275 0.279 0.283 0.288 0.292 0.374 0.378 0.382 0.386 0.390 0.394 0.398 0.402 0.406 0.410 0.415 0.419 0.423 0.427 0.431 0.442 0.446 0.450 0.454 0.458 0.463 0.467 0.471 0.475 0.479 0.483 0.491 0.496 0.500 0.504 0.508 0.512 0.520 0.525 0.529 0.533 0.537 0.541 0.545 0.549 0.553 0.558 0.562 0.566 0.574 0.578 0.582 0.586 0.591 0.698 0.702 0.707 0.711 0.715 0.719 0.727 0.731 0.018 0.024 0.028 0.032 0.110 0.115 0.130 0.134 0.138 0.158 0.162 0.166 0.170 0.174 0.179 0.183 0.187 0.191 0.195 0.200 0.204 0.208 0.212 0.216 0.220 0.224 0.228 0.233 0.237 0.296 0.300 0.304 0.309 0.313 0.317 0.321 0.325 0.329 0.333 0.338 0.342 0.346 0.0 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 13. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 15-17 November 1994, part I. The rotational phases are indicated to the right of each profile. a bi-dimensional paraboloid with linear and quadratic terms given by: aΩ2 eq + bΩeqdΩ + cdΩ2 + dΩeq + edΩ (2) The five coefficients are then used to solve for Ωeq and dΩ, and their errors. As discussed by Donati et al. (2003), the errors are c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? determined by computing the curvature radii of the χ2 paraboloid at its minimum and the correlation coefficient parameter between the two differential rotation parameters. 14 S.V. Jeffers, J.-F. Donati, A.Collier Cameron 1.0 c I / I 0.5 0.350 0.354 0.358 0.363 0.367 0.377 0.381 0.385 0.390 0.394 0.398 0.402 0.406 0.411 0.415 0.419 0.423 0.427 0.431 0.435 0.439 0.455 0.459 0.464 0.468 0.472 0.476 0.480 0.455 0.459 0.464 0.468 0.472 0.476 0.480 0.484 0.488 0.492 0.497 0.501 0.505 0.509 0.513 0.517 0.521 0.525 0.604 0.608 0.612 0.616 0.620 0.625 0.629 0.633 0.637 0.641 0.645 0.649 0.654 0.658 0.662 0.666 0.670 0.675 0.679 0.928 0.939 0.947 0.952 0.956 0.960 0.964 0.968 0.973 0.977 0.981 0.985 0.989 0.993 0.071 0.077 0.085 0.094 0.098 0.102 0.107 0.111 0.115 0.119 0.123 0.127 0.131 0.135 0.140 0.144 0.148 0.152 0.156 0.160 0.164 0.290 0.294 0.298 0.302 0.306 0.310 0.314 0.319 0.323 0.327 0.331 0.335 0.339 0.343 0.348 0.352 0.356 0.360 0.368 0.394 0.398 0.402 0.406 0.410 0.414 0.419 0.423 0.427 0.431 0.435 0.439 0.443 0.447 0.530 0.534 0.538 0.542 0.547 0.551 0.555 0.0 -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] -100 0 0 100 v [km/s] Figure 14. Maximum entropy fits (dashed line) to the LSD profiles (solid line) for 15-17 November 1994, part II. The rotational phases are indicated to the right of each profile. 5.1 Results For each epoch that comprises more than one night, we converged the imaging code to the fixed spot filling factors obtained when reconstructing the surface brightness images and computed a grid of Ω and dΩ models. The resulting χ2 landscape is shown in Fig- ure 16 for epoch 1992.05 and fitting the paraboloid (equation 2) to the data resulted in the values shown in Table 4. The results for epoch 1988.96 resulted in a map without any minimum, which is at- tributed to the poor quality of this data set. The temporal evolution of differential rotation over the epochs of this analysis is plotted in Figure 17. This plot clearly shows that the magnitude of tempo- c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 15 1 0.8 0.6 0.4 0.2 0 All Latitudes Latitudes 0o to 50o 90 80 70 60 50 40 30 20 10 0 0 0.003 0.001 0.002 Frac Spot Coverage e g a r e v o C t o p S . c a r F 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 Figure 15. Maximum Entropy surface brightness distribution for November 1994 (epoch 1994.87), where the vertical tics at the top of the plot indicate the phase coverage. The plot to the left shows the fractional spot coverage per latitude bin with integrated over longitude, while the plot below shows the fractional spot coverage per rotational phase bin, integrated over latitude. Epoch Ω eq dΩ cos2θs Ω s n (mrad d−1) (mrad d−1) (rad d−1) 1992.05 1993.89 1994.87 12 238.3 ± 2.1 12 249.5 ± 3.5 12 243.1 ± 3.1 60.1 ± 5.5 71.1 ± 9.3 73.6 ± 9.2 0.319 0.338 0.300 12.219 12.226 12.221 37761 26645 17141 Table 4. Summary of differential rotation parameters measured for AB Dor at each epoch, where Ωeq is the derived equatorial rotation rate at the 1σ 68% confidence interval, dΩ is the equator:pole differential rotation rate, column 4 is the inverse slope of the ellipsoid in the Ωeq-Ω plane (equal to cos2θs ref. Donati et al. (2003)), Ωs is the rotation rate at colatitude θs, and n is the total number of data points used in the imaging process. ral evolution is greater than the error bars of the differential rotation measurements, with the most noticeable change in dΩ of 11.0 mrad d−1 being between epochs 1992.05 and 1993.89. 6 DISCUSSION 6.1 Starspot distributions AB Dor is one of the most extensively Doppler imaged stars. In this analysis we have extended the work of Donati & Collier Cameron (1997); Donati et al. (1999) and Donati et al. (2003) to include epochs 1988.96, 1992.05, 1992.95, 1993.89 and 1994.87 to have a complete data set that has been processed using the same pro- cessing methods. One outstanding feature present at all epochs is a long-lived polar cap that is slightly fragmented at epoch 1988.96 and weaker at epoch 1992.95. Direct evidence for the presence of a polar cap has been shown by Jeffers et al. (2005) and Jeffers et al. (2006) using data from the Hubble Space Telescope of the RS CVn c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? binary SV Cam. The reconstructed surface images of this analysis also show many small mid to low latitude spots and a spot coverage that ranges from 6.25% at epoch 1992.95 to 9.2% at epoch 1994.87. These values bear no resemblance to the results of long-term photo- metric measurements, which show a brightness minimum in 1988 (Amado et al. 2001). The difference between the two results can be accounted for by the presence of chromopheric emission, which for young stars such as AB Dor, shows an anti-correlation with the star's photometric variations (Radick et al. 1998). For 1988.96 the reconstructed weak and fragmented polar cap is not conclusive due to the poor signal-to-noise of the data. How- ever, there is additional evidence from the results of Kurster et al. (1994) that AB Dor did not posses a polar cap in 1989. This could show that these large polar caps could disappear periodically. If such behaviour is cyclic then this could be related to a change in the dynamo nature of the star. The weaker polar cap for epoch 1992.95 probably results from the lower overall spot coverage at this epoch. The long-term stability of high, mid and low latitudes is shown 16 S.V. Jeffers, J.-F. Donati, A.Collier Cameron Figure 16. Reduced χ2 maps showing the differential rotation profile for epoch 1992.05. Only points that are within 1% of the minimum χ2 are shown. The central black region corresponds to points where the minimum value of differential rotation was obtained, while the outer grey points cor- respond to the 1.5 σ confidence ellipse (taking each parameter separately) on the differential rotation parameter plain. to vary on a timescale that is shorter than the temporal spacing of our observations. The evolution of small-scale magnetic features has been shown by Barnes et al. (2001) to be on a time-scale of less than one month for the active G2 dwarf He 699. The fractional spot coverage per latitude bin for each epoch of this analysis is shown in Figure 18. The relative fractional spot coverage for low- latitude features (between 0◦ and 50◦) is 24% for 1988.96, 17% for 1992.05, 26% for 1992.95, 30% for 1993.89 and 19% for 1994.87. The yearly distributions as shown in Figure 18 show that the peak does not become broader with time indicating that there is no ap- parent migration of high latitude spots towards the equator. How- ever this plot does show a global evolution of the spot distribution with possibly the polar spot becoming weaker with time and more spots forming at lower latitudes (compare plots for 1993.89 and 1998.96/1992.05). In contrast to these results, Jarvinen et al. (2005) show, from photometric data of AB Dor at similar epochs, that the mean spot latitude of 47-51◦ for 1992.05 and 45◦ for 1993.89 is significantly lower than the values of this analysis. The integrated brightness distributions are plotted as a func- tion of longitude for each surface brightness image, as shown in the lower panel of Figures 4, 7, 8, 12, and 15. These distributions show a variation of spot coverage with longitude, and regions of concentrated spot coverage that could be indicative of regions of enhanced magnetic activity or 'active longitudes'. Active longi- tudes are shown to be present on AB Dor by Jarvinen et al. (2005), where they reconstruct surface images comprising a primary and secondary spots from photometric data. The effect of high spot cov- erage can in, certain cases, result in spurious 'active longitudes' though these are always located at the quadrature points (Jeffers 2005). As previously discussed we show that the Doppler images re- constructed here are in broad agreement with the location of the primary spot of Jarvinen et al. (2005), but not with the existence or location of a secondary spot. The difference between the results of this analysis and those of Jarvinen et al. (2005) is not a result of inaccurate photometric data, but due to that it is not possible Figure 17. Differential rotation parameters measured for AB Dor as tabu- lated in Table 4 for epochs 1992.05, 1993.89 and 1994.87 (from this work) and 1995.94, 1996.99, 1999.00, 2000.93, 2001.99 (from Donati et al. 2003). The central point that is indicated by ◦ is for epoch 1992.05 from which a dashed line connects the points in chronological order. The central point is the best-fitting differential rotation measurement. For each differential rota- tion measurement, the 68% confidence ellipse is also plotted. to correlate phases of maximum and minimum photometric bright- ness with regions of the highest and lowest density of low latitude features. This is because the shape of the lightcurve can also be strongly influenced by the presence of high-latitude features and an off-centred polar cap. The effect of including high latitude spot fea- tures and polar cap is shown in the plots of the fractional spot cov- erage as a function of longitude (lower plot of Figures 4, 7, 8, 12, and 15) where plots are shown for spot features integrated over all latitudes and between 0◦ and 50◦. For epochs 1988.96 and 1993.89 there is little difference in the general shape of the two integrated spot coverage distributions. However, for epoch 1992.05 there are notable differences at phases 0.27 to 0.4 and 0.8 to 0.95 : in par- ticular the degree of fractional spottedness of the plot integrated over all latitudes (approximately 0.02 for phases 0.27 to 0.4 and 0.06 for phases 0.8 to 0.95), both show the same level of spotted- ness when integrated only between latitudes 0◦ to 50◦ when this is clearly not the case when integrated between 0◦ and 90◦. The same traits can be seen in the fractional spot distributions of epochs 1992.95 and 1994.87. Similar conclusions have also been reached by Donati et al. (2003) and Vogt et al. (1999). Additionally, our im- age reconstructions also do not show evidence for the migration of both primary and secondary spots at a constant separation or 'flip- flops' as shown in the work of Jarvinen et al. (2005). c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? Magnetic activity and differential rotation on AB Dor 17 e g a r e v o C t o p S l a n o i t c a r F 0.002 0.001 0 -0.001 -0.002 -0.003 -0.004 -0.005 1988 01/1992 12/1992 1993 1994 0 10 20 30 40 50 60 70 80 90 Latitude (degrees) Figure 18. Variation of the fractional spot coverage per latitude bin over the epochs of this analysis. 6.2 Differential rotation Surface differential rotation was measured for epochs with suf- ficient overlapping phase coverage, e.g. 1992.05, 1993.89 and 1994.87. In a complimentary paper, Donati et al. (2003) measure differential rotation on AB Dor for epochs after 1995 using the same method of this analysis. The resulting values of dΩ and Ωeq for all epochs are shown in Figure 17. This plot clearly shows that the results of this paper are generally higher than previous results with 1994.87 having the highest differential rotation ever measured on AB Dor. The smaller size of the error ellipses shown in Fig- ure 17 for epochs after 1994.87 are because the data set was taken over a longer time period enabling a more accurate differential ro- tation measurement to be made. The results so far, from 1992.05 to 2001.99, show no evidence for any cyclic behaviour. Another single dwarf with a similar spectral type to AB Dor and for which there have been multiple measurements of differen- tial rotation is the K2V dwarf LQ Hya. Only two measurements, a year apart, have been made with more than a magnitude of variation in ∆Ω ranging from 0.1942 rad d−1 to 0.01440 rad d−1. The large difference in measurements could result from the rotational period of LQ Hya being about three times that of AB Dor and hence the spatial resolution at the stellar surface is much less than that on AB Dor. This could result in the measurement of differential rotation being influenced by the short-term evolution of unresolved spots. The temporal evolution of differential rotation of AB Dor has also been determined by Collier Cameron & Donati (2002), us- ing a method that tracks individual starspots in the dynamic spec- trum, for the same epochs of this analysis. The results of this method show a stronger differential rotation measurement for all epochs; 1992.05 (Ωeq=12.2514 ± 0.0029 rad d−1 and dΩ=91.05 ± 13.19 mrad d−1), 1993.97 (Ωeq=12.2502 ± 0.0024 rad d−1 and dΩ=88.49 ± 7.47 mrad d−1) and 1994.87 (Ωeq=12.2481 ± 0.0047 rad d−1 and dΩ=66.84 ± 14.93 mrad d−1). Donati et al. (2003) discusses the discrepancy of measurements using the two methods. c(cid:13) 2006 RAS, MNRAS 000, 1 -- ?? It is concluded that there are two contributing factors; (i) the value of vsini used. The spot tracking method, which is sensitive to vsini, uses a value of 89 km s−1, while we use 91 km s−1 in this analy- sis, and (ii) the weighting of individual spots. The sheared imaged method of this analysis places a higher weighting on larger spots, while the spot tracking method places equal weighting on all spots. However, despite these small scale differences, the general trend of higher values of differential rotation for the epochs of this analysis are in agreement. Additionally, Donati et al. (2003) also measure differential ro- tation using magnetic features, which give a different result than using cool spots alone. This is interpreted as being evidence that the dynamo is distributed throughout the convective zone and not confined at the base. They also compare their results with models of the differential rotation in the convection zone and show that the internal rotation velocity field is not like that of the Sun, but more like that of rapid rotators where the angular velocity is constant along cylinders aligned with the rotation axis. Donati et al. (2003) surmise that changes in differential rotation could result from un- derlying dynamo processes. The temporal evolution of differential rotation will also have important consequences for the stellar structure. Large variations will alter the spherical oblateness of the star, such that if AB Dor was in a close binary system it would produce long-term changes in the star's orbital period. This is not applicable to the AB Dor A/C system given their comparatively large (2.3 AU) separation. Further support is given to the conjecture of Donati et al. (2003) by Applegate (1992), Lanza et al. (1998), Lanza & Rodon`o (1999) and Lanza (2005), where theoretical interpretations of the orbital period modulation in RS CVns is related to the operation of a hy- dromagnetic dynamo in the magnetically active star. The model of Applegate (1992) assumes that these periodic modulations are caused by the stellar magnetic cycle converting kinetic energy in the convective zone into large-scale magnetic fields. This is ex- tended by Lanza et al. (1998), and Lanza & Rodon`o (1999) to in- clude the effect of magnetic fields on the hydrostatic equilibrium of the magnetically active component. These models have been fur- ther extended by Lanza (2005) to include an improved treatment of angular momentum transport in the stellar convective zone, but they conclude that the method of Applegate (1992) is not suffi- cient to fully explain the mechanisms of orbital migration in close binaries. 7 CONCLUSIONS In this paper, Doppler images of the magnetically active star AB Dor show that its starspot coverage is dominated by a long-lived and stable polar cap and variable high to low latitude spot cover- age. The exception to this is the surface brightness reconstruction of epoch 1988.96 where there is evidence of a weak and fragmented polar cap. There is no cyclic behaviour found in either the latitude distribution of spots or the spot coverage fractions. Our surface brightness reconstructions generally show longitudes where there is a concentration of spots. However, we do not find a second or minor spot concentration which would verify the presence of ac- tive longitudes or flip-flop cycles on AB Dor. We have made the first measurements of differential rotation on AB Dor for epochs 1992.05, 1993.89 and 1994.87. The results show a temporal evolution, with epoch 1994.87 showing the high- est value of differential rotation ever measured on AB Dor. To a first order approximation the temporal evolution of differential rotation 18 S.V. Jeffers, J.-F. Donati, A.Collier Cameron show the same variation as the results for the same data recon- structed by Collier Cameron & Donati (2002). The results of this work when combined with other previously published papers repre- sents the first long-term analysis of the detailed temporal evolution of differential rotation on magnetically active stars. ACKNOWLEDGEMENTS This paper is based on observations made using the 3.9 m Anglo- Australian Telescope, the 3.6 m telescope at ESO and the 4 m telescope at CTIO. We thank the referee Steve Saar for suggest- ing several improvements to the paper. SVJ currently acknowl- edges support from a personal Marie Curie Intra-European fellow- ship funded within the 6th European Community Framework Pro- gramme. While at St Andrews University SVJ was supported by PPARC and a scholarship from the University of St Andrews, and would like to thank the Scottish International Education Trust for financing a Travel Grant for a collaborative visit to OMP. REFERENCES Amado P. J., Cutispoto G., Lanza A. F., Rodon`o M., 2001, in ASP Conf. Ser. 223: 11th Cambridge Workshop on Cool Stars, Stellar Systems and the Sun AB Doradus: Long and Short Term Light Variations and Spot Parameters (CD-ROM Directory: con- tribs/amado). pp 895 -- 900 Applegate J. H., 1992, ApJ, 385, 621 Babcock H. W., 1965, in ASSL Vol. 3: Plasma Space Science Vol. 1, The solar magnetic cycle. p. 7 Barnes J. R., Cameron A. C., Donati J.-F., James D. J., Marsden S. C., Petit P., 2005, MNRAS, 357, L1 Barnes J. R., Collier Cameron A., James D. J., Steeghs D., 2001, MNRAS, 326, 1057 Brown S. F., Donati J.-F., Rees D. E., Semel M., 1991, A&A, 250, 463 Brun A. S., Toomre J., 2002, ApJ, 570, 865 Bruning D. H., 1981, ApJ, 248, 274 Close L. M., Lenzen R., Guirado J. C., Nielsen E. L., Mamajek E. E., Brandner W., Hartung M., Lidman C., Biller B., 2005, Nat, 433, 286 Collier Cameron A., 1995, MNRAS, 275, 534 Collier Cameron A., Donati J.-F., 2002, MNRAS, 329, L23 Collier Cameron A., Duncan D. K., Ehrenfreund P., Foing B. H., Kuntz K. D., Penston M. V., Robinson R. D., Soderblom D. R., 1990, MNRAS, 247, 415 Collier Cameron A., Unruh Y. C., 1994, MNRAS, 269, 814 Collier Cameron A., Walter F. M., Vilhu O., Bohm T., Catala C., Char S., Clarke F. J., Felenbok P., Foing B. H., Ghosh K. K., Hao J., Huang L., Jackson D. A., Janot-Pacheco E., Jiang S., Lagrange A.-M., Suntzeff N., Zhai D. S., 1999, MNRAS, 308, 493 Donati J.-F., Brown S. F., 1997, A&A, 326, 1135 Donati J.-F., Cameron A. C., Semel M., Hussain G. A. J., Petit P., Carter B. D., Marsden S. C., Mengel M., L´opez Ariste A., Jeffers S. V., Rees D. E., 2003, MNRAS, 345, 1145 Donati J.-F., Collier Cameron A., 1997, MNRAS, 291, 1 Donati J.-F., Collier Cameron A., Hussain G., Semel M., 1999, MNRAS, 302, 437 Donati J.-F., Collier Cameron A., Petit P., 2003, MNRAS, 345, 1187 Donati J.-F., Mengel M., Carter B., Marsden S., Collier Cameron A., Wichmann R., 2000, MNRAS, 316, 699 Gray D. F., 1977, ApJ, 211, 198 Guirado J.C. et al. ., 1997, ApJ, 490, 835 Innis J. L., Thompson K., Coates D. W., Lloyd Evans T., 1988, MNRAS, 235, 1411 Jarvinen S. P., Berdyugina S. V., Tuominen I., Cutispoto G., Bos M., 2005, A&A, 432, 657 Jeffers S. V., 2005, MNRAS, 359, 729 Jeffers S. V., Barnes J. R., Cameron A. C., Donati J.-F., 2006, MNRAS, 366, 667 Jeffers S. V., Cameron A. C., Barnes J. R., Aufdenberg J. P., Hus- sain G. A. J., 2005, ApJ, 621, 425 Kurster M., Schmitt J. H. M. M., Cutispoto G., 1994, A&A, 289, 899 Kurucz R. L., 1993, CDROM # 13 (ATLAS9 atmospheric mod- els) and # 18 (ATLAS9 and SYNTHE routines, spectral line database), Cambridge, MA Lanza A. F., 2005, MNRAS, 364, 238 Lanza A. F., Catalano S., Cutispoto G., Pagano I., Rodono M., 1998, A&A, 332, 541 Lanza A. F., Rodon`o M., 1999, A&A, 349, 887 Leighton R. B., 1964, ApJ, 140, 1547 Leighton R. B., 1969, ApJ, 156, 1 Marsh T. R., 1989, PASP, 101, 1032 Martin T. J., Davey S. C., 1995, MNRAS, 275, 31 Parker E. N., 1955, ApJ, 122, 293 Petit P., Donati J.-F., Collier Cameron A., 2002, MNRAS, 334, 374 Rudiger G., Kuker M., 2002, A&A, 385, 308 Radick R. R., Lockwood G. W., Skiff B. A., Baliunas S. L., 1998, ApJS, 118, 239 Reiners A., Schmitt J. H. M. M., 2002, A&A, 384, 155 Unruh Y. C., Collier Cameron A., Cutispoto G., 1995, MNRAS, 277, 1145 Vilhu O., Gustafsson B., Walter F. M., 1991, A&A, 241, 167 Vogt S. S., Hatzes A. P., Misch A. A., Kurster M., 1999, ApJS, 121, 547 c(cid:13) 2006 RAS, MNRAS 000, 1 -- ??
astro-ph/0104087
2
0104
2001-05-23T13:00:43
The impact of bars on the mid-infrared dust emission of spiral galaxies: global and circumnuclear properties
[ "astro-ph" ]
We study the mid-infrared properties of a sample of 69 nearby spiral galaxies, selected to avoid Seyfert activity contributing a significant fraction of the central energetics, or strong tidal interaction, and to have normal infrared luminosities. These observations were obtained with ISOCAM, which provides an angular resolution of the order of 10 arcsec (half-power diameter of the point spread function) and low-resolution spectro-imaging information. Between 5 and 18 microns, we mainly observe two dust phases, aromatic infrared bands and very small grains, both out of thermal equilibrium. On this sample, we show that the global F15/F7 colors of galaxies are very uniform, the only increase being found in early-type strongly barred galaxies, consistent with previous IRAS studies. The F15/F7 excesses are unambiguously due to galactic central regions where bar-induced starbursts occur. However, the existence of strongly barred early-type galaxies with normal circumnuclear colors indicates that the relationship between a distortion of the gravitational potential and a central starburst is not straightforward. As the physical processes at work in central regions are in principle identical in barred and unbarred galaxies, and since this is where the mid-infrared activity is mainly located, we investigate the mid-infrared circumnuclear properties of all the galaxies in our sample. We show how surface brightnesses and colors are related to both the available molecular gas content and the mean age of stellar populations contributing to dust heating. Therefore, the star formation history in galactic central regions can be constrained by their position in a color-surface brightness mid-infrared diagram.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: missing; you have not inserted them ASTRONOMY AND ASTROPHYSICS The impact of bars on the mid-infrared dust emission of spiral galaxies: global and circumnuclear properties⋆ H. Roussel1, M. Sauvage1, L. Vigroux1, A. Bosma2, C. Bonoli3, P. Gallais1, T. Hawarden4, S. Madden1, and P. Mazzei3 1 DAPNIA/Service d'Astrophysique, CEA/Saclay, 91191 Gif-sur-Yvette cedex, France 2 Observatoire de Marseille, 2 Place Le Verrier, 13248 Marseille cedex 4, France 3 Osservatorio Astronomico di Padova, 5 Vicolo dell'Osservatorio, 35122 Padova, Italy 4 Joint Astronomy Center, 660 N. A'ohoku Place, Hilo, Hawaii 96720, USA Received 14 December 2000 / Accepted 22 March 2001 Abstract. We study the mid-infrared properties of a sample of 69 nearby spiral galaxies, selected to avoid Seyfert activity contributing a significant fraction of the central energetics, or strong tidal interaction, and to have normal infrared luminosities. These observations were ob- tained with ISOCAM, which provides an angular resolu- tion of the order of 10′′ (half-power diameter of the point spread function) and low-resolution spectro-imaging infor- mation. Between 5 and 18 µm, we mainly observe two dust phases, aromatic infrared bands and very small grains, both out of thermal equilibrium. On this sample, we show that the global F15/F7 colors of galaxies are very uni- form, the only increase being found in early-type strongly barred galaxies, consistent with previous IRAS studies. The F15/F7 excesses are unambiguously due to galactic central regions where bar-induced starbursts occur. How- ever, the existence of strongly barred early-type galaxies with normal circumnuclear colors indicates that the rela- tionship between a distortion of the gravitational potential and a central starburst is not straightforward. As the physical processes at work in central regions are in principle identical in barred and unbarred galaxies, and since this is where the mid-infrared activity is mainly lo- cated, we investigate the mid-infrared circumnuclear prop- erties of all the galaxies in our sample. We show how sur- face brightnesses and colors are related to both the avail- able molecular gas content and the mean age of stellar populations contributing to dust heating. Therefore, the star formation history in galactic central regions can be constrained by their position in a color-surface brightness mid-infrared diagram. Send offprint requests to: H. Roussel (e-mail: [email protected]) ⋆ Based on observations with ISO, an ESA project with in- struments funded by ESA Member States (especially the PI countries: France, Germany, the Netherlands and the United Kingdom) and with the participation of ISAS and NASA. Key words: galaxies: spiral -- galaxies: barred -- galax- ies: ISM -- stars: formation -- infrared: ISM: continuum -- infrared: ISM: lines and bands 1. Introduction As the high frequency of bars in galaxies becomes more evident (e.g. Eskridge et al. 2000), and as new techniques emerge to both observationally quantify their strength (Seigar & James 1998; Buta & Block 2001) and numer- ically simulate them, their effects on their host galaxies are of major interest, and in particular, it is worth check- ing whether they are indeed very efficient systems to drive nuclear starbursts in spiral galaxies. Numerous studies have dealt with the respective star formation properties of barred and non-barred spirals, mostly in the infrared, since this is the wavelength regime where starbursts are expected to be most easily de- tectable. Yet conclusions derived from such studies ap- pear to contradict each other, partly because the different selection criteria result in samples with a more or less pro- nounced bias toward starburst objects. For instance, in the IR-bright sample analyzed by Hawarden et al. (1986), an important fraction of SB and SAB galaxies (respectively strongly barred and weakly barred spirals in the classifica- tion of de Vaucouleurs et al. 1991) shows a 25 µm emission excess (with respect to 12 and 100 µm) absent in the SA subsample (non-barred spirals), which can be accounted for by a highly increased contribution of Galactic-like HII regions to the total emission. On the other hand, Isobe & Feigelson (1992), using a volume-limited sample and performing a survival analysis to take into account the frequent IRAS non-detections, found that the far-IR to blue flux ratio (FFIR/FB) is rather independent of the bar class. The contradiction is marginal since FFIR/FB does not give a direct estimation of the star formation activity, especially when dealing with quiescent normal galaxies: the blue light originates partly from young stars and, as 2 H. Roussel et al.: Impact of bars on MIR dust emission of spirals Isobe & Feigelson (1992) emphasize, FFIR/FB depends on the amount and spatial distribution of dust with respect to stars. The relationship between the 25 µm excess, quan- tified by F25/F12, and FFIR/FB in a galaxy sample with good quality data is indeed highly dispersed. Huang et al. (1996) investigated the 25 µm excess as a function of IR brightness and reconciled the two previous analyzes: a significant excess can occur only if FFIR/FB is larger than a threshold value of ≃ 0.3 . Therefore, a statistical effect of bars on star formation can be demonstrated only in suitably selected samples. Huang et al. (1996) also em- phasized that the difference between barred and unbarred spirals concerns only early types (S0/a to Sbc). Studies of the infrared excess in barred galaxies mostly rest on the integrated IRAS measurements, which do not allow the determination of the nature and location of regions responsible for this excess. However, dynam- ical models and observations at other wavelengths give evidence that the infrared activity should be concen- trated in circumnuclear regions (see for instance the study of NGC 5383 by Sheth et al. 2000). In addition, high- resolution ground-based observations near 10 µm of galaxy centers (Devereux 1987; Telesco et al. 1993) have shown that the dust emission is more concentrated in barred galaxies. Theoretically, bars are known to be responsible for large-scale redistribution of gas through galactic disks. In a strong barred perturbation of the gravitational poten- tial, shocks develop along the rotation-leading side of the bar and are associated with strong shear, as shown by Athanassoula (1992) and references therein (also Friedli & Benz 1993). They induce an increase of gas density which is traced by the thin dust lanes widely observed in bars, producing a contrasting absorption of optical light (Prendergast 1962, unpublished; Huntley et al. 1978). Due to these shocks, gas loses angular momentum and flows towards the circumnuclear region. This picture is con- firmed by direct observations of inward velocity gradients across bars in ionized gas lines, CO and HI (e.g. Lindblad et al. 1996; Reynaud & Downes 1998; Mundell & Shone 1999). Regan et al. (1997) derive a gas accretion rate of ≈ 1 M⊙ yr−1 into the circumnuclear ring of NGC 1530. Statistical evidence is also found for higher gas con- centrations in the center of barred galaxies (Sakamoto et al. 1999, who however observed only SABs, except NGC 1530), and for more frequent circumnuclear star- bursts in barred galaxies, as reported by Heckman (1980), Hawarden et al. (1986), Arsenault (1989) (who, more ex- actly, found more probable starbursts in galaxies with both bar and inner ring, supposed to be a signature of one or two inner Lindblad resonance(s)), Huang et al. (1996), Martinet & Friedli (1997) and Bonatto et al. (1998). Aguerri (1999) has moreover reported that the global star formation intensity of isolated spirals (mostly of late types) is correlated with bar strength as quantified by means of its projected axial ratio, which is surpris- ing in view of the very different timescales of bar evolu- tion (≈ 1 Gyr) and star formation in kpc-scale regions (≈ 107−8 yr). Indeed, Martinet & Friedli (1997), using carefully selected late-type galaxies, found no such cor- relation, the bar strength being quantified either by its deprojected axis ratio or its deprojected length relative to the disk diameter. The fact that only a fraction of strongly barred galaxies exhibit star formation excess (as evidenced by their IRAS colors) is explained by these au- thors with numerical simulations of bar evolution includ- ing gas physics. They show that a strong starburst occurs shortly after bar formation and quickly fades away (in typ- ically less than 1 Gyr); meanwhile, the strength and other properties of the bar evolve, but the bar remains strong if it was initially strong. The existence of strongly barred galaxies in a quiescent state is thus to be expected, pre- sumably because the available gas supply has been con- sumed in previous bursts. This paper is aimed at characterizing the mid-infrared excess in barred galaxies, with the possibility to carry out a detailed and systematic spatial analysis due to the good angular resolution of ISOCAM (the half-power beam di- ameter is less than 10′′ at 7 µm), and hence to locate unambiguously sites of enhanced infrared activity. Al- though dust is a more indirect tracer of young stars than far-ultraviolet ionizing radiation or optical recombination lines, the infrared emission suffers relatively minor extinc- tion effects, which are very difficult to correct and hamper shorter wavelength studies. In a companion paper (Rous- sel et al. 2001a, hereafter Paper II), we have shown that in galactic disks, mid-infrared emission is a reliable star for- mation indicator. Here, we concentrate on central regions of galaxies where the dust heating regime is markedly dif- ferent from that in disks. For this purpose, we have analyzed a sample of 69 nearby spiral galaxies, imaged at 7 and 15 µm with the camera ISOCAM on board ISO (described by Cesarsky et al. 1996c). We have also obtained low-resolution spec- troscopic information for a few galaxies, enabling us to identify and separate the various dust components emit- ting between 5 and 18 µm. 7 µm images and F15/F7 flux density ratios of selected regions, together with optical im- ages, are presented in Roussel et al. (2001b) (hereafter the Atlas). For a description of data reduction and analysis, and a summary of morphological properties of the sample, the reader is also referred to the Atlas. 2. The galaxy sample The sample is intended to be representative of normal quiescent spirals, and contains galaxies of moderate in- frared luminosity. It covers three guaranteed time pro- grams of ISOCAM. The first one (Cambarre) consists of nearby barred galaxies, the second one (Camspir) of a few large-size spirals of special interest (NGC 1365, 4736, 5194, 5236, 5457 and 6744) and another subsample is H. Roussel et al.: Impact of bars on MIR dust emission of spirals 3 drawn from the Virgo cluster sample of Boselli et al. (1998) (Virgo program), containing relatively fainter and smaller galaxies, both barred and unbarred. This sample was sup- plemented by comparable spirals in the ISOCAM public archive, from the programs Sf glx (Dale et al. 2000) and Irgal (PI T. Onaka). All of the observations were reduced in the same way to form a homogeneous sample. The final set comprises 69 spiral galaxies at distances between 4 and 60 Mpc. We have divided them into three main categories according to morphological classes in the RC3 (de Vau- couleurs et al. 1991): SBs (accounting for about half the sample with 37 galaxies), SABs (20 galaxies) and SAs (12 galaxies). The latter two classes are merged to form the control sample to compare with SB galaxies. This sam- ple, although not statistically complete, has been selected according to the following requirements: -- All objects are relatively nearby, which ensures good spatial resolution with a 6′′ or 3′′ pixel size (the ex- tension of the central concentration is typically 5 -- 10 pixels in diameter). At the distances of the sample, a 3′′ pixel corresponds to linear sizes between 60 and 900 pc. Virgo galaxies, less extended and all imaged with 6′′ pixels in order to increase the signal to noise ratio, are resolved but with less detail. -- The sample was selected to avoid non-stellar activity as well as strong signs of tidal interaction, with a few exceptions in the Cambarre and Virgo subsamples, de- tailed in the following. -- Galaxies included in the first two programs are mod- erately inclined on the line of sight (i ≤ 50◦). This was not a requirement for the other programs, so that one third of the Virgo galaxies and one third of the supplementary galaxies are inclined by more than 60◦. -- The number of SA-SAB galaxies is comparable to that of SBs and both groups span the whole de Vaucouleurs spiral sequence from types S0/a to Sdm. -- Both barred and unbarred galaxies cover a large range of far-infrared luminosities (between 108.6 and 1011 L⊙bol), but none would be classified as an in- frared luminous galaxy, except NGC 7771 which is at the lower boundary of this class -- defined by 1011 < LFIR < 1012 L⊙bol (Sanders & Mirabel 1996). Another property which was not a selection criterion is that absolute blue magnitudes are equal to or greater than the typical magnitude of the Schechter luminosity function in the field, M ∗ B ≃ −21 (more exactly, they range between -21.17 and -17.38). Despite the incompleteness of the sample, we have checked that it is very similar to the magnitude-limited CfA galaxy sample (Thuan & Sauvage 1992), from the point of view of its infrared brightness normalized by blue starlight. For CfA spiral galaxies detected in all 4 IRAS bands and with blue magnitudes in the RC3, log (FFIR/FB) falls in the interval [−0.97; +0.98] with a mean value of 0.05. Using the same IRAS references as those in Thuan & Sauvage (1992), i.e. in order of prefer- ence Thuan & Sauvage (1992), Rice et al. (1988), Soifer et al. (1989) and Moshir et al. (1989), galaxies in our sample have log (FFIR/FB) in the interval [−1.58; +1.67] with a mean value of 0.01 . For that set of references, a Wilcoxon- Mann-Whitney (WMW) test indicates that the probabil- ity for the two populations to have the same FFIR/FB distribution is about 75%. We note that the IRAS 12 µm fluxes often disagree with our 7 and 15 µm fluxes, although the bandpasses overlap. Thus, when we use IRAS data, we take them from the references we consider the most reli- able (i.e. which provide the best match between 12 µm and our 7 -- 15 µm flux densities). In that case, log (FFIR/FB) falls in the interval [−0.77; +0.95] with a mean value of 0.02, and the WMW test gives a probability of about 40%. Hence, our sample is not different from optically complete samples regarding the fraction of the energy radiated in the infrared. Table 1 lists some general characteristics of the galax- ies. The morphological classification adopted is that of the RC3 (de Vaucouleurs et al. 1991). Although it is based on blue images, which may not be as appropriate as near- infrared images for detecting bars, many more galaxies are classified as barred in this catalog than for instance in Sandage & Bedke (1994). We have found only two galax- ies classified as SA in the RC3 and possessing a bar (as described in the following). A drawback of using the SB and SAB classes of the RC3 is that they do not consti- tute a measure of the bar dynamical strength. The bar strength is however difficult to quantify, and reliable mea- sures, such as those of Buta & Block (2001), are scarce. In the following, we will refer to bar lengths, normalized by the disk diameter, because longer bars are able to col- lect gas from inside a larger area and have low axis ratios, which are among the (unsatisfactory) quantities used to estimate bar strengths; bar lengths are in addition rela- tively easy to measure. The two sub-samples of spirals found in the field or loose groups and Virgo galaxies have been separated, be- cause they differ both in their aspect in the infrared (Virgo members are fainter and less extended) and in their en- vironment. Although Virgo is not a very rich cluster, the interaction of central galaxies with the intracluster gas and with their neighbours is likely to cause either a depletion or an enhancement of star formation activ- ity in the outer parts of disks and also to have global dynamical consequences. An extreme case is the galaxy NGC 4438 (= VCC 1043), whose very perturbed morpho- logical appearence was successfully modelled by Combes et al. (1988) as the result of a collision with NGC 4435. Several Virgo members have truncated HI disks due to the interaction with the cluster hot gas (Cayatte et al. 1990); a very clear example is NGC 4569 (= VCC 1690), which on optical photographs shows the juxtaposition of a bright and patchy inner disk structured by star forma- tion sites and dust lanes, and a low surface brightness and 4 H. Roussel et al.: Impact of bars on MIR dust emission of spirals very smooth outer disk with faint spiral arms. Severely HI-stripped galaxies can indeed be recognized in the op- tical as anemic (defined by van den Bergh 1976 as an in- termediate and parallel sequence between lenticulars and spirals), due to the suppression of star formation where the gas density is too low. Table 1 also indicates whether signatures of nuclear activity or tidal interaction exist. In addition to these, some galaxies deserve special com- ments (see the Atlas for more details) and should be con- sidered cautiously in the interpretation of the data set: -- NGC 337, 1385 and 4027 are strongly asymmetric and fit in the category of magellanic barred spirals. -- NGC 4691 and 1022 have an amorphous structure and highly centrally concentrated interstellar tracers. Their morphology is suggestive of merger results. Their star formation activity may therefore not be a consequence of the bar, since the latter was likely produced at the same time by the same cause, i.e. the merger. 3. Observations and photometric results All galaxies were observed with two broadband filters, LW3 (12 -- 18 µm) and LW2 (5 -- 8.5 µm), that we shall here- after designate by their central wavelength, respectively 15 and 7 µm. This was expected to provide F15/F7 col- ors directly linked with star formation intensity, since the LW2 filter covers the emission from a family of bands (see Sect. 4), which are ubiquitous in the interstellar medium, and LW3 was supposed to cover mainly a thermal con- tinuum observed to rise faster than the emission bands in star-forming regions, for instance from the IRAS F25/F12 ratio (Helou 1986); however, we will see that the picture is more complicated. Maps covering the whole infrared- emitting disk were constructed in raster mode. In all cases, the field of view is large enough to obtain a reliable de- termination of the background level, except for NGC 4736 and 6744. The pixel size is either 3′′ or 6′′, depending on the galaxy size. The half-power/half-maximum diameters of the point spread function are respectively 6.8′′/≃ 3.1′′ at 7 µm with a 3′′ pixel size, 9.5′′/5.7′′ at 7 µm with a 6′′ pixel size, 9.6′′/3.5′′ at 15 µm with a 3′′ pixel size and 14.2′′/6.1′′ at 15 µm with a 6′′ pixel size. The data reduc- tion is described in the Atlas. Since the emission from various dust species and atomic lines is mixed in the broadband filters (see Sect. 4), it is essential to complement our maps with spectro- imaging data. These allow an estimate of the relative im- portance of all species as a function of the location in- side a galaxy. We have thus obtained spectra between 5 and 16 µm of the inner disks (3′ × 3′ or 1.5′ × 1.5′) of five bright galaxies: NGC 613, 1097, 1365, 5194 and 5236 (Fig. 1). Spectra averaged over a few central pixels cover- ing approximately the extent of the circumnuclear region (left column) are compared with spectra averaged over the inner disk, excluding the central part and a possible ghost image (middle column). The right column shows the observed spectrum of the faintest pixels, consisting of the zodiacal spectrum contaminated by emission features from the target galaxy, because the field of view never extends beyond the galactic disk. For this reason, we can- not measure exactly the level of the zodiacal foreground to remove. Instead, as explained in the Atlas, we first fit a reference zodiacal spectrum to the average spectrum of the faintest pixels (excluding the spectral regions where emission features appear). The upper limit to the zodiacal foreground is set by offsetting the fitted spectrum within the dispersion range, with the additional constraint that the corrected disk spectrum remains positive; the lower limit is symmetric to the upper limit with respect to the fit. This makes little difference for the nuclear spectra but it does for the disk spectra, although it does not affect the spectral shape. Note that due to the configuration of the instrument, two different filters are used for the short and long wavelength parts of the spectra, and that a small offset can result at the junction of these filters, around 9.2 µm. The mid-infrared maps generally show an intense cir- cumnuclear source. Decomposing surface brightness pro- files into a central condensation and a disk (see details in the Atlas), we define a radius for this circumnuclear re- gion, RCNR. Total fluxes and fluxes inside RCNR are listed in Table 2 with the background level for each broadband filter. Explanations about the method employed for pho- tometry and the estimation and meaning of errors can be found in the Atlas. The dominant uncertainty arises from memory effects for relatively bright galaxies, and from other sources of error (essentially the readout and photon noise) for faint galaxies, especially at 15 µm. For galaxies drawn from the Sf-glx project, the number of ex- posures per sky position is very small (≃ 10) and does not allow a proper estimate of memory effects: their pho- tometric errors are thus especially ill-determined. Typical errors are ≈ 10% at 7 µm and 18% at 15 µm. Note that flux density calibration uncertainties, which are of the order of 5 to 10%, are not included. However, this is a systematic effect, hence not affecting relative fluxes. 4. Nature of the mid-infrared emitting species The spectra shown in Fig. 1 are strikingly similar to one another. They contain some features also seen in spectra of reflection nebulae, atomic and molecular envelopes of HII regions, atmospheres of C-rich evolved stars as well as the diffuse interstellar medium. We can thus safely as- sume that the results obtained on these resolved Galac- tic objects can be readily extrapolated to the emission of galaxies where individual sources are no longer resolved. The emission between 5 and 16 µm is dominated by the so-called unidentified infrared bands (UIBs) at 6.2, 7.7, 8.6, 11.3 and 12.7 µm. Our spectra also display weak features which have previously been detected as broad fea- H. Roussel et al.: Impact of bars on MIR dust emission of spirals 5 tures in SWS spectra of starburst objects (Sturm et al. 2000) at e.g. 5.3, 5.7, 10.7, 12.0, 13.6, 14.3 and 15.7 µm1. A 7.0 µm feature can tentatively be identified as an [ArII] line (6.99 µm) or an H2 rotational line (6.91 µm), but our spectral resolution (∆λ/λ ≈ 40) prevents a more defi- nite identification. We note however that the [ArII] line has been identified in the high-resolution SWS spectra of starburst galaxies (Sturm et al. 2000). It was originally proposed by Duley & Williams (1981) that UIBs are due to organic functional groups on car- bonaceous grains. L´eger & Puget (1984) instead favoured vibration modes of C-C and C-H bonds only, in large poly- cyclic aromatic molecules not in thermal equilibrium with the local radiation field (the so-called PAH model). The constancy of the spectral energy distribution of UIBs, re- gardless of the radiation field (Sellgren 1984; Uchida et al. 2000), implies an impulsive heating mechanism, where upon absorption of a single UV photon, the carriers un- dergo a very rapid and large temperature increase and then radiatively cool before the next absorption. Alter- native candidates for the UIB carriers are various hydro- genated and oxygenated carbon grains, amorphous but partially ordered at the smallest scale (Borghesi et al. 1987; Sakata et al. 1987; Papoular et al. 1989), much simi- lar to the idea of Duley & Williams (1981). Recent work by Boulanger et al. (1998b) indicates that UIBs are not due to molecules such as PAHs, but more likely to aggregates of several hundred atoms. In the interstellar medium surrounding the OB associ- ation Trapezium (Roche et al. 1989), the Orion bar (Giard et al. 1994) and M17 (Cesarsky et al. 1996a; Tran 1998), these features are detected in the HII region and the molec- ular cloud front (provided projection effects are minor), but the emission peaks at the photodissociation interface (see also Brooks et al. 2000). UIB carriers are likely de- stroyed in HII region cores, although the estimation of the critical radiation field necessary to obtain a significant re- duction in UIB carrier abundance still remains to be done (compare e.g. Boulanger et al. 1988, 1998a; Contursi et al. 2000). 1 We also detect a weak and unknown emission feature be- tween 9.3 and 9.9 µm, which seems brighter, relatively to UIBs, in disks than in central regions. However, the very poor signal to noise ratio of disk spectra does not allow us to be conclusive. It cannot be an artefact due to the change of filter since that change occurs after the feature is observed. It is too narrow to be emitted by silicates. The identification with ionized PAHs (see e.g. Allamandola et al. 1999) would be inconsistent with the fact that the flux ratio of this feature to classical UIBs seems higher in regions of low radiation density and excitation than in central regions. It is also unlikely that it corresponds to the H2 rotational line at 9.66 µm since this is characteristic of warm and excited molecular clouds in starburst nuclei (e.g. Spoon et al. 2000). Finally, we mention that it also matches in wavelength a feature from the CH3 functional group at 9.6 µm (Duley & Williams 1981). While the 7 µm flux in spiral galaxies essentially con- sists of the UIB emission, the 15 µm filter covers the emis- sion from mainly two dust species: the hot tail of a contin- uum attributed to very small grains (VSGs) of the order of 0.5 -- 10 nm in size and most often impulsively heated like UIB carriers (D´esert et al. 1990), and also UIBs. The red wing of the 11.3 µm band contributes little, but the band at 12.7 µm and the emission plateau that connects it to the 11.3 µm band can be important; the smaller UIB features listed above also contribute, although to a lesser extent. When spatial resolution is high enough, the emis- sion from VSGs and UIB carriers can be clearly separated: around M 17 and in the reflection nebula NGC 7023, the VSG continuum strongly peaks in a layer closer to the ex- citation sources than the UIBs, inside the ionized region for M 17 (Cesarsky et al. 1996a, 1996b). Therefore, the F15/F7 flux ratio decreases with increasing distance from the exciting stars of an HII region. In the spectra of all five galaxies (Fig. 1), the intensity ratios of UIBs are remarkably stable, which is a common property of a variety of astronomical sources (Cohen et al. 1986; Uchida et al. 2000). The only highly varying fea- ture is the VSG continuum that is best seen longward of 13 µm. It has various amplitudes and spectral slopes in galactic nuclei. It remains very modest compared to that in starburst galaxies (Tran 1998; Sturm et al. 2000), and is hardly present in averaged disks. In Paper II, we show that the integrated mid-infrared luminosity of normal spi- ral disks is dominated by the contribution from photodis- sociation regions (where the UIB emission is maximum). From a comparison with Hα luminosities, we show that this predominance of the photodissociation region emis- sion results in the fact that, when integrated over the disk, the UIB emission is a good tracer of massive young stars. Finally, as alluded to earlier, a number of fine-structure lines can be present in the mid-infrared spectral range, although their contribution to the broadband flux is al- ways negligible in spirals. In normal galaxies, the most prominent is the [NeII] line at 12.81 µm, which at the spectral resolution of ISOCAM is blended with the UIB at 12.7 µm. No lines from high excitation ions such as [NeIII] at 15.56 µm are convincingly detected, and the [NeII] line at 12.81 µm is weak, since the intensity of the blend with the UIB at 12.7 µm, relative to the isolated UIB at 11.3 µm, is rather stable in different excitation conditions. Some variation however exists. To compare the strength of the [NeII] line in our galaxies to that observed by Forster-Schreiber et al. (2001) in the starburst galax- ies M 82, NGC 253 and NGC 1808, we have measured in a similar way the flux of the blend F12.75 above the pseudo- continuum drawn as a straight line between 12.31 and 13.23 µm, and the flux of the 11.3 µm UIB F11.3 with its respective continuum level defined in the same way be- tween 10.84 and 11.79 µm. We find that the energy ratio F12.75/F11.3 of circumnuclear regions decreases from 0.67 in NGC 1365 to 0.60 in NGC 613 and 5236, 0.52 in 6 H. Roussel et al.: Impact of bars on MIR dust emission of spirals Fig. 1. Spectra of central regions (left) and the inner disk (middle). The upper and lower limits are determined from limits on the zodiacal spectrum shown with dotted lines (right), adjusted using the average spectrum of the faintest pixels, also shown with its dispersion. The flux unit for all spectra is mJy arcsec−2. H. Roussel et al.: Impact of bars on MIR dust emission of spirals 7 NGC 1097 and 0.47 in NGC 5194; in the averaged inner disks of NGC 1365, 5236 and 5194, where it is still mea- surable, it takes the approximate values 0.5, 0.45 and 0.4 . These figures are much lower than those observed in cores of starburst galaxies by Forster-Schreiber et al. (2001), where it can reach 1.7, and argue for a generally small contribution of [NeII] to the spectra. Adopting as the in- trinsic F12.7/F11.3 UIB energy ratio the minimum value of F12.75/F11.3 that we measure in our spectra, i.e. 0.4, we obtain a maximum [NeII] equivalent width of 0.22 µm in the nucleus of NGC 5236. As for the UIBs, their equivalent widths in disks and central regions range respectively be- tween EW(12.7)= 0.3 -- 0.6 µm and EW(11.3)= 1.2 -- 1.9 µm (these numbers do not take into account broad UIB wings that occur if the bands are described by Lorentzians). Our estimates give circumnuclear values for F12.7/F[NeII] be- tween 1.5 and 5.7, that we can compare with the results of Sturm et al. (2000) in the starburst galaxies M 82 and NGC 253 from their ISOSWS spectra with a high spec- tral resolution (λ/∆λ ≈ 1500, versus ≈ 40 for ISOCAM). They obtain values of 0.96 and 1.32 . The contribution from the [NeII] line to our spectra is thus confirmed to be negligible with respect to starburst galaxies. 5. The dependence of F15/F7 on spiral and bar types and comparison with IRAS results In their infrared analysis, Huang et al. (1996) pointed out that bars are able to significantly enhance the total star formation only in early-type galaxies (mixing all types between S0/a and Sbc). It is also known that bars do not share the same properties all through the Hubble se- quence: among early types, or more exactly in spirals with large bulges, since the relationship between Hubble type and bulge to disk ratio is far from direct (e.g. Sandage & Bedke 1994; Seigar & James 1998), they tend to be longer (Athanassoula & Martinet 1980; Martin 1995), and their amplitude, with respect to that of the underlying axisym- metric potential, tends to be higher. For instance, Seigar & James (1998), using K band photometry to trace the stellar mass, find that galaxies with the strongest bars have bulge to disk mass ratios between 0.3 and 0.5 . For larger bulges, their number of galaxies is too low to derive any meaningful bar strength distribution. Early-type bars host little star formation, except near their ends and at their center, whereas late-type galaxies generally harbor HII regions all along the bar (Garc´ıa-Barreto et al. 1996), which suggests that their shocks are not as strong as in early types (Tubbs 1982). Inner Lindblad resonances be- tween the gas and the density wave, which appear when there is sufficient central mass concentration and when the bar rotates more slowly than (Ω − κ/2)max (where Ω is the gas circular rotation frequency and κ the epicyclic frequency), and which presence induces straight and off- set shocks along the bar (Athanassoula 1992), are also typically expected in early-type galaxies. These structural Fig. 2. (a): Integrated mid-infrared color F15/F7 as a function of morphological type for unbarred or weakly barred galaxies, represented respectively by open circles and crossed circles. Virgo galaxies are identified by their VCC number (see Table 1) and others by their NGC num- ber. (b): Same as (a) for strongly barred galaxies. differences have consequences on the efficiency of bars to drive massive inward gas flows. We show in Fig. 2a the distribution of F15/F7 accord- ing to morphological type (as given in the RC3) for the control subsample including only SA and SAB galaxies. For this population -- excepting NGC 41022 -- , the mid- 2 This galaxy shows a peculiar structure, with a central lens- like body or fat oval of moderate length (Dbar/D25 ≈ 0.2) sur- rounded by an external pseudo-ring probably associated with a Lindblad resonance. NGC 4102 is thus a genuine weakly barred galaxy, and not a SB. However, such a dynamical structure is still efficient to drive inward mass transfer. We also point to the very strong concentration of its mid-infrared emission (see Ta- ble 2), a property common to early-type barred spirals (Fig. 6). 8 H. Roussel et al.: Impact of bars on MIR dust emission of spirals infrared color is remarkably constant around a value of 1 (ranging from 0.7 to 1.2). This is rather typical of the color of the surface of molecular clouds exposed to radiation fields ranging from that observed in the solar neighbor- hood to that found in the vicinity of star-forming regions. F15/F7 colors observed toward HII regions are typically of the order of 10, while those of photodissociation re- gions range between 2 and the HII region values (Tran 1998). The fact that F15/F7 remains of the order of 1 in most galaxies -- it also shows generally little variation from pixel to pixel in disks -- indicates that, at our angular resolution, emission from HII regions and their immediate surroundings is diluted by the larger neighboring interstel- lar medium (at a mean distance of 20 Mpc, 3′′ represent 300 pc). In fact, in the Atlas, we show that even in giant star-forming complexes that can be identified in the maps, F15/F7 rarely exceeds 2 -- 3. The case of strongly barred spirals is more complex (Fig. 2b): whereas many of them share the same integrated colors as their unbarred counterparts, an important frac- tion shows a color excess, the maximum color being above 2.5 instead of 1.2 for SA(B)s. Furthermore, such an ex- cess occurs only among the earliest morphological types, from SB0/a to SBb. Note that in bulges, the envelopes of K-M stars can contribute an important fraction of the mid-infrared emission. However, this would be negligible at 15 µm and mostly affect the 7 µm band: correcting for such an effect would only re-inforce the observed trend. We also qualify that observation by noting that two galax- ies, NGC 1022 and NGC 4691, have likely experienced a merger; gas may therefore have sunk to the center as a result of the violent energy dissipation in the merger, and not simply under the influence of the bar, which actually may have been formed during the interaction. Dismissing these two objects however does not change the fact that the color distribution of the strongly barred galaxies shows 15 µm excesses that are absent from that of weakly barred or unbarred spirals. One can wonder whether cluster galaxies introduce a bias in our sample, because a number of them are per- turbed by their environment and thus may have an un- certain morphological type. Koopmann & Kenney (1998) have shown that a significant fraction of early-type spirals in Virgo have been "misclassified" due to their dearth of star formation in the disk. The degree of resolution of spi- ral arms into star formation complexes is indeed one of the three criteria defining the Hubble sequence, but it is not unambiguously linked to the bulge to disk ratio. Concern- ing several Virgo members of our sample, the bulge is very small for the attributed type (Sandage & Bedke 1994), in such cases defined mostly by the disk appearance. This is of course related to the anemia phenomenon, due to gas deficiency caused by interaction with the intracluster medium. Of our Virgo galaxies of types S0/a -- Sb, 10/14 are HI-deficient, versus 3/9 for types Sbc -- Sdm (see Ta- ble 1, where def > 1.2 has been adopted as the criterion for HI deficiency). This apparent segregation with mor- phological type certainly results from the above classifica- tion bias. Thus, differentiating galaxies in Fig. 2 according to their true bulge to disk ratio would cause an under- representation of SA-SAB early-type spirals, which make the crucial part of our comparison sample. If we had to discard completely the early-type SA-SAB subsample, the maximum allowed conclusion from Fig.2 would be that we observe a color excess in a fraction of early-type strongly barred galaxies, without excluding the possibility of such an excess in early-type non-barred galaxies, in which case another mechanism for mass transfer would have to be thought of. However, at least five early-type SA-SAB spirals re- main which are not HI-deficient and thus unlikely to suf- fer from the above bias, namely VCC 92 = NGC 4192, NGC 3705, NGC 4736, NGC 5937 and NGC 6824. We do not consider NGC 3885, SA0/a in the RC3, because it looks like a genuine barred galaxy: its bulge is elongated in a direction distinct from the major axis of outer isophotes and crossed by dust lanes; it is furthermore classified as such by Vorontsov-Velyaminov & Arkhipova (1968) and Corwin et al. (1985). These galaxies show no global color excess, like the rest of the SA-SAB subsample, and like a number of bona-fide early-type SB galaxies with normal HI content. Hence, our view should not be too strongly dis- torted by the classification bias. We have also checked the influence of this bias on HI-deficient barred galaxies with a color excess. On optical images, VCC 836 = NGC 4388 unambiguously resembles classical early-type spirals, with a prominent bulge crossed by thick dust lanes; VCC 460 = NGC 4293 stands between HI-deficient and HI-normal galaxies, and also has an early-type aspect. The case of VCC 1326 = NGC 4491 is not that clear, because it is a low-mass galaxy. Our sample thus confirms and extends to the ISOCAM bands a phenomenon that was evidenced from IRAS ob- servations by Hawarden et al. (1986) and Huang et al. (1996), namely that a significant fraction of SB galaxies can show an excess of 25 µm emission (normalized to the emission at 12 or 100 µm) compared with SA and SAB galaxies. The case of SAB galaxies is in fact unclear: some of them show such an excess according to Hawarden et al. (1986), but they are indistinguishable from SAs in the analysis of Huang et al. (1996). From the present ISO- CAM data, it already appears that indeed SA and SAB galaxies share similar mid-infrared properties. In order to compare more directly our results with IRAS-based results, Fig. 3 shows the relationship between F15/F7 and F25/F12. For log (F25/F12) < 0.3 (which is close to the value given by Hawarden et al. (1986) as the limit for the presence of a 25 µm excess), F15/F7 shows no systematic variation and SA, SAB and SB galaxies are well mixed. Above that threshold, SB galaxies strongly domi- nate (the SAB galaxy with high colors is NGC 4102) and the F15/F7 ratio follows the increase of F25/F12. Given the H. Roussel et al.: Impact of bars on MIR dust emission of spirals 9 though not statistically significant, between SBs with no 15 µm excess and SA-SAB galaxies: the LFIR/LB logarith- mic means and dispersion factors are 1.1 and 2.7 for SBs with no excess, and 0.8 and 2.2 for SAs-SABs. That mid-infrared color excesses occur only in SB galaxies indicates that somehow, a global increase of the interstellar radiation field intensity is linked to the pres- ence of a strong bar, although this condition is clearly not sufficient. The fact that many barred galaxies earlier than SBb appear very similar in their integrated color to their unbarred counterparts means that no simple link exists between the bar class, the bulge-to-disk ratio and the on- set of a starburst in normal spirals. Several intervening parameters can be thought of: the true strength of the bar in dynamical terms (the separation into SB and SAB classes is subjective and too rough, and in a recent study, Buta & Block 2001 show that the SB class includes a wide range of actual bar strengths); the available gas content inside corotation; the star formation efficiency along bars and in central regions; the timescales for starburst acti- vation and exhaustion; interaction with a companion or with the intracluster gas. Some of these effects can be in- vestigated in the present sample. We will discuss them in Sect. 7, but first we turn our attention to mid-infrared properties of the central regions, as defined in Sect. 3 and in the Atlas. 6. The role of central regions As the presence of a bar is expected to influence the star formation in the circumnuclear region and much less in the disk (except in the zone swept by the bar), we are naturally led to emphasize the relative properties of nuclei and disks. Maps shown in the Atlas demonstrate that cen- tral regions, observed in the infrared, are prominent and clearly distinct from other structures, much more than on optical images. In Fig. 4 we plot the fraction of the total 15 µm flux originating from the central region (inside the radius RCNR) as a function of the global F15/F7 color. Galax- ies for which a central region could not be defined on the mid-infrared brightness profiles are also shown, and are attributed a null central fraction. Galaxies are not dis- tributed at random in this plot, but rather on a two-arm sequence that can be described in the following way: (1) high F15/F7 colors are found exclusively in systems where a high fraction of the flux is produced in the circumnuclear regions; (2) galaxies with small F15/F7 ratios (< 1.2) are found with all kinds of nuclear contributions. The bar class appears to play a part in the location of galaxies in this diagram, although this is not clear-cut: all galaxies with high circumnuclear contribution (> 40%) and large F15/F7 colors (> 1.2) are SB galaxies, apart from NGC 4102, while SA-SAB galaxies are quite indis- tinguishable from one another and cluster in the small nuclear contribution (< 30%) and low F15/F7 color cor- Fig. 3. Comparison of mid-infrared colors from ISO (F15/F7) and IRAS (F25/F12). F25 always contains the VSG emission (see Sect. 4) whereas at low temperatures, F15 is dominated by UIBs, which explains the constancy of F15/F7 below a threshold of F25/F12 ≃ 2. The same convention as in Fig. 2 applies for the representation of SA, SAB and SB classes. We have indicated the names of the Sy2 galaxy NGC 4388 = VCC 836 (see Sect. 7.1), and of the two galaxies with the lowest F15/F7 colors (note that NGC 6744 was not entirely mapped and that its in- tegrated color is likely a lower limit). nature of dust components whose emission is covered by the 7 µm to 25 µm filters (Sect. 4), this behavior can be explained as follows. The classical interpretation for the variation of F25/F12 is that it increases with the radia- tion field due to the stronger contribution of VSGs to the 25 µm than to the 12 µm emission, which collects mostly UIB emission (D´esert et al. 1990; Helou 1986). The fact that the F15/F7 ratio remains insensitive to the variation of F25/F12 for log (F25/F12) < 0.3 implies that in this regime, VSGs provide little flux to both ISOCAM bands as well. Past this threshold, the increase of F15/F7 signals that the VSG continuum has entered the 15 µm bandpass and contributes an ever increasing fraction. The galaxies with a 15 µm excess (F15/F7 above 1.2, or 0.08 dex) also distinguish themselves from the rest of our sample by having on average larger far-infrared to blue luminosity ratios. For this subsample, LFIR/LB spans the range [0.6; 7.3] with a logarithmic mean of 2.2 and disper- sion by a factor 2.2, while LFIR/LB of the complementary subsample falls in the interval [0.2; 9.0], has a logarithmic mean of 0.9 and dispersion by a factor 2.4 . However, the 15 µm-excess galaxies have far-infrared luminosities that are equivalent to those observed in the rest of the sample. Hence, in these galaxies with a VSG emission excess, a higher fraction of the total emission is reprocessed in the whole infrared range. There is also a slight difference, al- 10 H. Roussel et al.: Impact of bars on MIR dust emission of spirals ner of the graph. There is also a clear preponderance of SB galaxies in all the centrally dominated range. Only two SA-SAB galaxies show very high concentration fractions, NGC 3885 and NGC 4102. The latter galaxy was already discussed; for NGC 3885, strong indications exist that its bar class is incorrect (see the discussion in Sect. 5). However, it is quite significant that SB galaxies cover both sequences in Fig. 4 and in particular are found all through the sequence of varying flux concentration and low F15/F7 color. Therefore, Fig. 4 shows that high global F15/F7 colors require that the flux concentration be high, and that the galaxy be SB, but none of these two proper- ties is enough to predict that the global F15/F7 ratio will be high. To understand the importance of the flux concen- tration, let us first study separately the colors of central regions and those of disks. Fig. 5 compares the F15/F7 distributions observed in the disk and in the central regions of our galaxies (when- ever the radius of the central regions RCNR, fitted on 7 µm brightness profiles, could not be defined, the galaxy has been considered as a pure disk). These histograms indicate that F15/F7 ratios of circumnuclear regions are higher than those of disks (and this is a systematic prop- erty, verified for each individual galaxy except NGC 4736 and 6744, whose central regions are dominated by old stellar populations). Colors of disks are fairly constant and close to the integrated colors of SA-SAB galaxies (F15/F7 = 0.89 ± 0.14 for the 1σ dispersion), whereas cir- cumnuclear colors form a broader distribution extending towards high values (F15/F7 = 1.59 ± 0.78). The cause for this difference of colors can easily be seen in the spectra of Fig. 1: in all spectra with sufficient signal- Fig. 4. Relationship between the central flux fraction at 15 µm and the integrated F15/F7 color. Galaxies with no identifiable central regions, i.e. surface brightness profiles consistent with a single disk component, have been placed at a null ordinate. Fig. 5. Compared histograms of F15/F7 colors averaged in disks and in circumnuclear regions. The galaxies used are respectively those whose disk is not strongly contami- nated by the central component (the excluded galaxies are NGC 1022, NGC 4691, VCC 1419 = NGC 4506, NGC 1326 and NGC 3885), and those with central regions that could be adjusted on surface brightness profiles (otherwise the galaxy is considered to be composed only of a disk). The isolated galaxy with a very low central color is NGC 6744, which is clearly devoid of young stars all inside its inner ring. to-noise ratio, the relative intensities of the UIBs are al- most unchanged from galaxy to galaxy, or from central regions to disks. On the contrary, the level and spectral slope of the continuum seen longward of 13 µm is highly variable and always stronger in the central regions than in the disks. This continuum is attributed to VSGs (see Sect. 3) and its presence in the 15 µm band is a charac- teristic sign of intense star formation (e.g. Laurent et al. 2000). The reason why high global F15/F7 colors require a high flux concentration can be directly derived from Fig. 5: only the central regions of galaxies are able to reach high F15/F7 colors, and they have to dominate the integrated emission to affect the global color. Furthermore, the fact that the two color histograms overlap explains why a high flux concentration does not necessarily imply a high F15/F7 color. We however still have to identify the property or prop- erties required, in addition to belonging to the SB class, for a galaxy to show a high mid-infrared flux concentration. We have seen in Fig. 2 that the morphological type plays a major part in the presence of high colors. Fig. 6 shows the evolution of the concentration fraction as a function of morphological type. It confirms that for SB galaxies, there is a definite trend for the central flux fraction to rise as the morphological type gets earlier. More precisely, SB H. Roussel et al.: Impact of bars on MIR dust emission of spirals 11 disk-dominated galaxies. This suggests that they host at their center larger concentrations of gas and dust than in the average of galaxies of the same Hubble type, but for some yet undetermined reasons, they presently undergo smooth star formation instead of a nuclear starburst. We propose that either the net gas inflow rate to the center has decreased (due to a slower replenishment from the inner disk which would have been previously partially de- pleted in gas, or a smaller efficiency of the evolved bar to make gas lose its angular momentum) or, since star for- mation bursts occur on a much shorter timescale than bar life, that we are imaging these objects at a period of qui- escence in-between bursts. Concerning this last point, see the results of the simulations of Martinet & Friedli (1997) and the population synthesis estimates of Kotilainen et al. (2000) for the circumnuclear rings of NGC 1097 and 6574. 7. Origin of the circumnuclear infrared excess 7.1. Non-stellar activity Of the SB galaxies, four are known to host a Seyfert nucleus: in order of decreasing flux fraction from the central condensation, VCC 836 = NGC 4388, NGC 1365, NGC 1097 and NGC 1433. For these, the high central color could arise from dust heated by non-stellar radiation from the accretion disk and halo of the central object and would thus not necessarily indicate the presence of massive stars. For NGC 1097, we have the direct visual evidence that the contribution from the active nucleus to the circumnuclear emission is negligible, since the central mid-infrared source is resolved into the well-known star-forming ring, which is very bright, and a faint point source at the nucleus. Cor- recting the images of NGC 1097 for dilution effects with a procedure analog to CLEAN (see the Atlas for more de- tail), we obtain fractions of the total circumnuclear fluxes contributed by the nuclear point source of less than 3% at 7 µm and about 1% at 15 µm. This central source was measured inside a radius of 3′′, while the ring extends be- tween radii ≈ 6′′ and 12′′. We can also inspect the low-resolution spectra between 5 and 16 µm of the central regions of NGC 1097 and 1365 (left column of Fig. 1). Indeed, Genzel et al. (1998) and Laurent et al. (2000) have shown that a strong contin- uum at 5 µm and small equivalent widths of the UIBs are signatures of dust heated by an active nucleus. Yet all our spectra are similar to that of the inner plateau of NGC 5194 (≈ 50′′ in diameter) -- which also contains a weak Seyfert nucleus, but completely negligible -- and to that of NGC 5236: they are dominated by UIBs in the 5 -- 10 µm range and the underlying continuum at 5 µm is comparatively very low. We conclude that in these galax- ies, the contribution of non-stellar heating to the emission observed inside RCNR is small. The cases of NGC 4388 and NGC 1433 can only be discussed on the basis of imaging results. The central con- densation of NGC 1433 is large (we have determined a Fig. 6. Fraction of total 15 µm fluxes arising from the cen- tral condensation, as a function of morphological type. As in Fig. 4, galaxies with no identifiable central regions have been placed at a null ordinate. The central fraction of F7 fluxes, not shown here, has a very similar behavior, with only slightly lower values. galaxies with central fractions greater than 40% are found predominantly among galaxies earlier than Sb. It is less clear in Fig. 6 whether SA-SAB galaxies fol- low a similar or a different trend, partly because of the lack of such bar classes in our sample for types S0/a and Sa, and also because types Sab and Sb may be incorrect due to the morphological classification bias affecting clus- ter galaxies, as already discussed in Sect. 5. In that section however, we emphasized the existence of a set of five early- type SA-SAB spirals which do not suffer from morpho- logical misclassification: VCC 92 = NGC 4192, NGC 3705, NGC 4736, NGC 5937 and NGC 6824. As apparent in Fig. 6, they all have a low central flux fraction, much lower than that observed in SB galaxies in the same range of types. This supports the view that the trend seen for in- creasing concentration fraction with earlier type concerns only SB galaxies (or peculiar objects like VCC 1043), SA- SAB galaxies having a generally low concentration factor whatever their type. We can summarize our findings in this section in the following way: integrated F15/F7 colors of galaxies are generally of the order of 1. However, F15/F7 is often higher in central regions. Spiral galaxies with high F15/F7 colors must simultaneously be (1) dominated by their central re- gions, (2) of bar type SB, and (3) of morphological type earlier than Sb. However, the reverse is not true: as can be seen in Fig. 6, NGC 5383 (a Markarian galaxy), 1672, 1365 and 1097 for instance fulfill these conditions -- between 55 and 75% of their 15 µm radiation comes from small central regions (respectively 17, 8, 6 and 8% of the optical diam- eter) -- yet their F15/F7 color is very similar to that of 12 H. Roussel et al.: Impact of bars on MIR dust emission of spirals diameter of 31′′ ≈ 1.7 kpc) and extremely smooth, much flatter than the point spread function: we therefore con- sider unlikely a major contribution from the LINER/Seyfert nucleus, which should manifest itself as a point source. For NGC 4388, we cannot conclude and the active nucleus may be dominant. We can only mention that its global color is lower than that of VCC 1326 = NGC 4491, and this is marginally true as well for the nucleus, and that the nu- cleus of VCC 1326 is not classified as active3. Hence, the presence of Seyfert nuclei does not mod- ify our interpretation that high mid-infrared colors in the present sample are not due to dust heated by non-stellar photons and should rather signal the existence of central starbursts. 7.2. Circumnuclear starbursts We now examine the most likely cause of the 15 µm emis- sion excesses detected in our sample, central starbursts triggered by the bar dynamical effects. We warn that NGC 1022 and NGC 4691 should be considered apart: their dust emission comes almost exclusively from cen- tral regions of ≈ 1 kpc, but this is more likely due to a past merger than to the influence of the bar, which may have been formed or transformed simultaneously as the starburst event was triggered. 7.2.1. Available molecular gas To see if a significant difference exists between the central molecular gas content of circumnuclear starburst galax- ies and quiescent ones, we have searched the literature for single-dish CO(1-0) data in the smallest possible beams. Single-dish data are better suited to our purpose than interferometric data since the latter are scarcer and do not collect all the emission from extended structures. The conversion of CO antenna temperatures to molecular gas masses is approximate for two main reasons: the H2 mass to CO luminosity ratio varies with metallicity and physi- cal conditions; and the derivation of CO fluxes requires the knowledge of the source structure, because it is coupled to the antenna beam to produce the observed quantity which is the antenna temperature. Sensible constraints on the structure of CO emission can be drawn from that observed in the mid-infrared. As dust is physically associated with gas, the mid-infrared emission spatial distribution should follow closely that of the gas, but be modified by the distribution of the 3 Far-infrared diagnostics of nuclear activity are ambiguous for NGC 4388: its infrared to radio flux ratio as defined by Con- don et al. (1991) is q = 2.27, and its spectral index between 25 and 60 µm (de Grijp et al. 1987) is α = 1.23. Both values in- dicate that stellar and non-stellar excitations may contribute in comparable amounts to the infrared energy output. Con- cerning VCC 1326, its spectral index between 25 and 60 µm, α = 1.67, is rather typical of starbursts. star-forming regions that provide the heating, and which are likely more concentrated than the gas reservoir. Since gaussian profiles provide an acceptable description of most infrared central regions at our angular resolution, we have therefore assumed that the CO emitting regions are of gaussian shape, with half-power beam width (HPBW) be- tween one and two times that at 7 µm. The 7 µm HPBW were derived by matching gaussian profiles convolved with the point spread function to the observed 7 µm profiles4. To find the meaning of various antenna temperatures (with various corrections) and which conventions are used in the literature, the explanations of Kutner & Ulich (1981) and Downes (1989) were of much help. We con- R scale5. We then at- verted given temperatures to the T ∗ tempted a correction of antenna to source coupling, as- suming a gaussian source and a gaussian diffraction pat- tern with angular standard deviations θS and θB. The relationship below follows for the source brightness tem- perature Tb, which is averaged over the beam in the ob- servation, whereas we want to recover its intrinsic value over the source extent: R × (θ2 T ∗ S + θ2 B) = Tb × θ2 S. Table 1 contains the beam width of the observa- tions and the derived H2 masses for the adopted ref- erences. A conversion factor f = N (H2) / I(CO) = 2.3 1024 molecules m−2 (K km s−1)−1 (Strong et al. 1988) has been used to compute the mass as: MH2 / (2 mH) = f × Z Tb dV (K km s−1) × 2 π (θS D)2 × (1 − exp(− 1 2 (αCNR / θS)2)) where mH is the hydrogen atom mass and D the distance (in m). In the above formula, we estimate the mass only inside the angular radius αCNR used for the infrared pho- tometry of circumnuclear regions. Only when the central regions are resolved and mapped is there no need to as- sume a brightness distribution. H2 masses derived in this way are probably not more precise than by a factor three, including the dispersion of the factor f , but the dynamic range in the sample is still sufficient to allow a discussion of the results. 4 In NGC 1530, the scales of the molecular gas and infrared concentrations are of the same order. In NGC 1022, the source HPBW is estimated to be ≈ 17′′, which is ≈ 2.9 times that of the innermost infrared regions. However, these are clearly more extended than a central gaussian and not representative of normal CNRs, since likely gathered by a merger. For this and other galaxies whose central regions are clearly structured (i.e. NGC 1097 and NGC 4691), detailed CO maps where the source is resolved were used. It is also the case for NGC 1530, 5236 and 6946. 5 which includes corrections for atmospheric attenuation and all instrumental effects except antenna to source coupling. H. Roussel et al.: Impact of bars on MIR dust emission of spirals 13 brightness of the central regions. This is expected, since the amount of dust scales with that of gas, which essen- tially consists of the molecular phase in central regions of galaxies. More interesting is Fig. 7b where we show the evolution of the F15/F7 color inside RCNR as a function of the same quantity as in Fig. 7a. For the majority of our sample, F15/F7 tends to rise, within a very large disper- sion, when the molecular gas mean density increases (it roughly doubles when the H2 surface brightness varies by 1.2 dex). However, a few galaxies dramatically depart from this trend: for colors higher than 2.5 (log F15/F7 > 0.4), there is a reversal in the sense that hot circumnuclear re- gions seem to be depleted in molecular gas, with respect to the normal H2 content -- color distribution. Although one can think of several reasons why their molecular content may be underestimated (the standard conversion factor may not apply for these galaxies due to their starburst nature or possibly due to a lower metal- licity), it is unlikely that this is the case. First, the im- plied underestimation factors appear quite large, at least 4 to 10. Second, if we were to correct the H2 masses by these factors to bring the galaxies within the trend ob- served in Fig. 7b, then these objects would become ab- normal in Fig. 7a, with a deficit of 7 µm emission6. The four deviating galaxies do not share a common property which would make them special with respect to all the others. NGC 4519 and IC 1953 are similar SBd galaxies, VCC 1326 = NGC 4491 is a small and low-luminosity SBa, and VCC 836 = NGC 4388 is an edge-on Seyfert SBab (for which the molecular content may be ill-determined due to the integration of the CO line throughout the disk). We thus propose the following interpretation for the galaxies that wander off the main trend in Fig. 7b: the main distribution corresponds to galaxies where the cen- tral starburst is more and more intense, as indicated by the high gas surface densities and colors. Galaxies at the turnover of the sequence may be observed in a phase of their starburst (not necessarily common to all galaxies) when it has consumed or dispersed most of the accumu- lated gas, because of a higher star formation efficiency. This suggests an interesting analogy with HII regions, for which the distinction between "ionization-bounded" and "density-bounded" is made (see Whitworth 1979, also for a discussion of the efficiency of molecular cloud dispersal by young stars). Dust should then be depleted too; how- ever, because of the presence of massive stars, the remain- ing dust is exposed to a very intense radiation field and reaches a high F15/F7 color. This ratio may also increase due to the fact that the dust which was mixed with rather dense molecular clouds, of low F15/F7 color, has been dis- persed too. Alternatively, the concentrations of molecular gas in these galaxies may be more compact than in the others and diluted in our large beam (we cannot exclude 6 Note that it is however conceivable that a fraction of UIB carriers are destroyed, which would cause such a deficit. Fig. 7. (a): 7 µm surface brightness as a function of the average H2 surface density, both inside the circumnuclear regions defined by mid-infrared photometry (CNR). The limits on H2 mass are not true error bars, but simply indi- cate the effect of varying the scale of the gaussian distribu- tion from once to twice that measured at 7 µm (see text). (b): F15/F7 color as a function of the average H2 surface density, both inside the CNR, with the same convention for error bars as in (a). Although the beam of CO observations is in general larger than αCNR, it remains (except for NGC 337) smaller than the diameter of the bar which collects gas from inside corotation, believed to be located close to the end of the bar (Athanassoula 1992), so that it is still meaningful to compare our measurements on infrared condensations to CO data. Fig. 7a shows the variation of the 7 µm surface bright- ness as a function of the average molecular gas surface density inside RCNR. Higher densities of the molecular material are associated with an increase in the infrared 14 H. Roussel et al.: Impact of bars on MIR dust emission of spirals that the mid-infrared distribution includes an unresolved core which dominates the color). A confirmation of the above scenario clearly requires better measurements of the central gas content and high-resolution characterization of the starbursts. Leaving the four galaxies in the upper left quadrant of Fig. 7b apart, the data support an interpretation in terms of starburst with standard properties: the infrared activ- ity in galactic centers can be stronger when the available molecular gas is denser. 7.2.2. Color of the central concentration and age of the starburst Fig. 8 indicates how the F15/F7 color inside RCNR varies with the 15 µm surface brightness in the same aperture. In principle, the mid-infrared surface brightness can in- crease either because the amount of dust in the considered area is higher (such as observed in Fig. 7a), or because the energy density available to heat the dust increases. The trend for higher F15/F7 ratios at large 15 µm surface brightnesses seen in Fig. 8 indicates that indeed, the in- crease of the 15 µm surface brightness is at least partly due to rise of the mean energy density in the CNR. In this diagram again, the galaxies with a peculiar behavior in Fig. 7b stand apart, well above the locus defined by the least absolute deviation fit7 (dashed line). This supports the fact that the trend seen in Fig. 7b is not due to an un- derestimation of the H2 content, and lends further credit to the interpretation presented in Sect. 7.2.1. Another study by Dale et al. (1999) has already dealt with the joint variations of mid-infrared surface bright- nesses and colors. However, contrary to Fig. 8 where the surface brightnesses and colors are those of the same phys- ical region (the CNR) in a large sample of galaxies, in the Dale et al. (1999) study, resolution elements inside the target galaxies are first binned according to their surface brightness before the mean color of the bin is computed. As a result, a bin does not correspond to a physical object. We simply note that if galactic central regions are binned by surface brightness in Fig. 8, then the obtained mean locus is comparable to those shown by Dale et al. (1999). The galaxies with the highest central F15/F7 colors (F15/F7 > 2.5) and which stray from the main trend are barred, but their bars are of moderate lengths (once deprojected and normalized by the optical diameter). In NGC 4519, NGC 4102, VCC 1326 = NGC 4491 and IC 1953, for which it could be estimated, Dbar/D25 ≈ 0.2 -- 0.3, when this ratio ranges between 0.06 and 0.67 in galax- ies with measurable bar length. This confirms that the central activity, signalled by a high F15/F7 color, is not 7 We give this name to a fit where the quantity to be min- imized is the sum of absolute values of the distances between the data points and the line. This method is less sensitive to outliers than the least squares fit. an increasing function of bar strength, as can be expected from the different timescales for star formation and bar evolution. Since the bar strength alone is not sufficient to ex- plain the observed mid-infrared colors, and since the ob- servational uncertainties are much smaller than the scat- ter present in Fig. 8, one may suspect that part of this scatter is due to intrinsic properties of each of the cir- cumnuclear starbursts considered. Indeed, given that mid- infrared emission likely traces star formation on timescales longer than, for instance, recombination lines, it is reason- able to expect that for similar mid-infrared brightnesses (corresponding to similar gas and energy densities), the mid-infrared color could vary as a function of the age of the stellar populations responsible for dust excitation. Since star formation does not happen instantaneously all through a ≈ 1 kpc region and likely occurs in cycles trig- gered by instabilities, these stellar populations are multi- ple and their ages should be weighted to reflect the suc- cessive generations of stars contributing to dust heating. Using the population synthesis results of Bonatto et al. (1998), based on ultraviolet spectra between 1200 and 3200 A, we can estimate the mean stellar age in the central 10′′ × 20′′, weighted by the fraction of luminosity emitted at 2650 A by different population bins. This was possi- ble for eleven galaxies of our sample in common with the Fig. 8. Variation of the mid-infrared color with the 15 µm surface brightness (in mJy arcsec−2), both inside the same aperture centered on circumnuclear regions. The mean er- ror bar is shown in the upper left corner. The average location of disks in the diagram in terms of average color and global surface brightness (the disk area being delim- ited by the blue isophote µB = 25 mag arcsec−2) would be at (−2.2, −0.05). The dashed line represents the for- mal least absolute deviation fit, performed including all the galactic central regions (used to define the "color de- viation" in Fig. 9). H. Roussel et al.: Impact of bars on MIR dust emission of spirals 15 sample of Bonatto et al. (1998). We compare in Fig. 9 this mean age to the F15/F7 color deviation, defined as the difference between the observed color and that pre- dicted by the mean distribution of all galaxies (indicated by the least absolute deviation fit in Fig. 8) at the same 15 µm surface brightness. For this purpose, we performed the mid-infrared photometry in a slit aperture identical to that used by Bonatto et al. (1998). We indeed see that the younger the weighted age, the higher the central F15/F7 color deviation. This is thus in agreement with our hy- pothesis that much of the color variations in Fig. 8 may be due to age variations of the exciting populations. There are grounds to think that some scatter in Fig. 9 is due to the methodology adopted by Bonatto et al. (1998) in their study. They have grouped galaxies of their sample according to spectral resemblance, morphological type and luminosity, and co-added all UV spectra of each group in order to increase the signal to noise ratio before performing the population synthesis. However, it may not be fully justified to average spectra of different galaxies with the same overall shape but different spectral signa- tures. A further drawback of this study is that it cannot properly take into account extinction, because of the lim- itation to a small spectral range in the UV: the derived very low extinctions are meaningless. That is why the two galaxies departing from the well-defined trend described above could owe their age to the method rather than to their intrinsic properties: -- VCC 1690 = NGC 4569 is assigned the same large weighted age as NGC 7552 (their UV spectra are co- added), but has a higher color excess with respect to the mean distribution in Fig. 8. In fact, Maoz et al. (1998) detected very strong P Cygni absorption lines of high-excitation ions (CIV, SiIV, NV) characteristic of the winds of massive young stars (< 6 Myr), and its spectrum between 1220 and 1590 A is nearly identical to that of the starburst NGC 1741B. -- NGC 1433 is co-added with NGC 4102, which results in a small weighted age. Yet its color excess is much lower than that of NGC 4102 and more comparable to that of NGC 7552. There is some evidence that the extinction in the central regions of NGC 1433 is much lower than in other galaxies: whereas available Balmer decrement measures indicate Hα absorptions of the order of 2 -- 3 in other nuclei, the decrement given by Diaz et al. (1985) for NGC 1433 indicates A(Hα) ≈ 0.9. Even if the Balmer decrement is not a good extinction mea- sure, it is instructive to compare values in different galaxies. We also notice that NGC 1433 is the only one among strongly barred spirals which has an amorphous circumnuclear region, with no hot spot that would in- dicate the presence of massive stellar clusters. To conclude, Fig. 8 is a strong indication that the mid- infrared emission in circumnuclear regions is influenced by successive episodes of star formation over relatively long Fig. 9. The abscissa indicates the mean age of stellar pop- ulations, according to the synthesis results of Bonatto et al. (1998), including the first six elements of their base, stellar clusters to which they attribute ages between 0 and 0.7 Gyr for the first five and in the interval 0.7-7 Gyr for the last one, but excluding the oldest element, an ellip- tical bulge representing ages between 7 and 17 Gyr. The ages are weighted by the fraction of the flux at 2650 A that each different population emits. The contribution from continuous star formation has been approximated by a constant flux fraction (equal to the minimum value) and subtracted, in order to consider only successions of bursts. The ordinate is the difference between the mea- sured F15/F7 color and the color expected from the mean relationship between central surface brightnesses and col- ors shown in Fig. 8. For this graph, the photometry was performed inside the same apertures as in Bonatto et al. (1998), 10′′ ×20′′, and error bars show the effect of varying the slit orientation, which is not given by Bonatto et al. (1998). periods of time: on the mean, the F15/F7 color is a sen- sitive function of the mid-infrared surface brightness, but this relationship is modulated by the mean age of the stel- lar populations. A strinking example of this is NGC 4736. Its central mid-infrared brightness is in the high range, but its central F15/F7 ratio is low, in accordance with its large mean stellar age confirmed by Taniguchi et al. (1996). From optical population synthesis, they find that a central starburst occurred about 1 Gyr ago in this galaxy, and that subsequent nuclear star formation has proceeded at a low rate. Combining this result with that of Sect. 7.2.1, we can form the following sketch of what determines the mid- infrared properties of circumnuclear regions: the central surface brightness is connected to the amount of gas, as expected if gas-to-dust ratios are relatively constant. How- ever, accumulation of gas in the center allows the trigger- 16 H. Roussel et al.: Impact of bars on MIR dust emission of spirals ing of intense star formation, so that the interstellar ra- diation field increases, reflected in higher F15/F7 ratios. Fig. 8 and 9 suggest then that deviations from this sim- ple description can be related to the star formation history of the circumnuclear regions. On-going starbursts produce excess F15/F7 colors, while faded starbursts are associated with F15/F7 deficits. Additional variation in mid-infrared colors may arise from differences in metallicity and in the compactness of the starburst, with consequences on the amount and na- ture of the dust, but this is out of the scope of the present study. 8. Summary and conclusions We have studied the mid-infrared activity induced by bars in a sample of 69 nearby spiral galaxies with infrared lumi- nosities spanning a large range below the class of luminous infrared galaxies. We have found that: -- The mid-infrared emission of the normal galaxies in our sample is essentially contributed by a thermal con- tinuum from very small grains (VSGs) longward of 10 µm and the family of aromatic bands (UIBs) de- tected in a wide diversity of environments. It is the variation of the VSG component with respect to the UIBs that is responsible for F15/F7 changes in our galaxies. From the comparison with observations of resolved Galactic regions, this can be related to an increase of the filling factor of star forming complexes by photoionized regions, hence a decrease of the contri- bution to the mid-infrared emission from neutral and molecular media. -- There is a dichotomy between spiral disks, where the integrated F15/F7 color is close to 1 and shows little dispersion, and circumnuclear regions, where F15/F7 ranges from disk-like to high values (up to 4). We have found no indication that destruction of UIB carriers occurs at the scale of circumnuclear regions, although it would be desirable to analyze infrared spectroscopic data of the most active galaxies, to elaborate on this. -- We confirm that barred spirals distinguish themselves from unbarred galaxies in the sense that they can reach higher F15/F7 colors. This effect is however re- stricted to early morphological types, in agreement with previous IRAS-based studies (Hawarden et al. 1986; Huang et al. 1996). We show unambiguously that this emission excess arises in circumnuclear re- gions which can completely dominate the mid-infrared emission, although their size remains modest (D CNR ranges between 2 and 26% of the optical diameter in the whole sample, with no clear dependency on Hubble type). Galaxies with a global color excess are all dom- inated by their central regions. This is a confirmation of predictions from hydrodynamical models (Athanas- soula 1992; Friedli & Benz 1993), according to which a barred perturbation, through tidal torques and shocks, induces substantial mass transfer towards circumnu- clear regions. We observe the consequences of these gas flows, i.e. the intense star formation that they fuel. -- An important fact to mention is that only a fraction of early-type barred galaxies can be distinguished from unbarred galaxies in their infrared properties. Several interrelated parameters may explain this quiescence of many barred galaxies. With the present data we are unable to address this issue and thus only list possible explanations: a bar evolves on a much longer timescale than a starburst (Martinet & Friedli 1997) and the ac- cretion rate by a bar is slow; the inward mass transfer is regulated by the depth of the potential well, the in- tensity of the shocks inside the bar, the star formation efficiency along the path of the inflowing gas before it reaches central regions (Martin & Friedli 1997), etc. Although the presence of a bar can be an efficient means of triggering circumnuclear starbursts, the dust emission processes in central regions are the same in barred and unbarred galaxies. We have studied the prop- erties of these central regions at the degree of detail acces- sible to our spatial resolution. Several physical properties were found to control the mid-infrared color F15/F7. -- The estimated molecular gas content inside the central regions (RCNR): as expected, the mid-infrared bright- ness at 7 µm tends to rise with increasing mean gas density, which reflects the physical association of dust with molecular material. The F15/F7 color is also cor- related with the mean gas density. As a higher gas con- tent allows more efficient star formation on relatively large scales, according to the Schmidt law and stabil- ity criteria or other empirical laws (Kennicutt 1998), this supports an interpretation of mid-infrared colors in terms of starburst intensity. A few galaxies, show- ing the most extreme F15/F7 ratios, depart from the general trend. This could be interpreted as a transient evolutionary state of the starburst during which most of the gas has been consumed or dispersed, although a definite assessment of the effect requires better molec- ular line data. -- The age of the stellar populations heating dust: dust is sensitive to star formation on relatively long timescales, as can be expected from the fact that it can be heated by optical and near-UV photons. How- ever, the strength of the VSG continuum is more sensi- tive to the radiation energy density and hardness than UIBs, and the results that we report here indicate that F15/F7 excesses are linked to the weighted age of the exciting stellar populations. Mid-infrared colors are therefore influenced by the previous star formation history over at least 1 Gyr, and depend on the fraction of the ultraviolet radiation power contributed by young populations created in a contemporary starburst, with H. Roussel et al.: Impact of bars on MIR dust emission of spirals 17 respect to intermediate-age populations from already faded bursts. Cesarsky C.J., Abergel A., Agnese P. et al., 1996c, A&A 315, L32 Acknowledgements. We thank our referee, Louis Martinet, for his helpful remarks. The ISOCAM data presented in this paper were analyzed using and adapting the CIA package, a joint development by the ESA Astrophysics Division and the ISOCAM Consortium (led by the PI C. Cesarsky, Direction des Sciences de la Mati`ere, C.E.A., France). References Aalto S., Booth R.S., Black J.H. & Johansson L.E.B., 1995, A&A 300, 369 (CO ref.: AB) Aaronson M., Huchra J., Mould J.R. et al., 1982, ApJS 50, 241 (HI ref.: AH) Aguerri J.A.L., 1999, A&A 351, 43 Allamandola L.J., Hudgins D.M. & Sandford S.A., 1999, ApJ 511, L115 Andreani P., Casoli F. & Gerin M., 1995, A&A 300, 43 (CO ref.: AC) Appleton P.N., Foster P.A. & Davies R.D., 1986, MNRAS 221, 393 (HI ref.: AF) Arsenault R., 1989, A&A 217, 66 Ashby M.L.N., Houck J.R. & Matthews K., 1995, ApJ 447, 545 (nuc. type ref.: A) Athanassoula E., 1992, MNRAS 259, 345 Athanassoula E. & Martinet L., 1980, A&A 87, L10 Bajaja E., Wielebinski R., Reuter H.P., Harnett J.I. & Hummel E., 1995, A&AS 114, 147 (CO ref.: BW) van den Bergh S., 1976, ApJ 206, 883 Bonatto C., Pastoriza M.G., Alloin D. & Bica E., 1998, A&A 334, 439 Borghesi A., Bussoletti E. & Colangeli L., 1987, ApJ 314, 422 Boselli A., Lequeux J., Sauvage M. et al., 1998, A&A 335, 53 Boselli A., Casoli F. & Lequeux J., 1995, A&AS 110, 521 (CO ref.: Bo) Boulanger F., Abergel A., Bernard J.P., et al., 1998a, in "Star formation with the Infrared Space Observatory", J. Yun & R. Liseau Eds, ASP Conf. Series 132, 15 Boulanger F., Boissel P., Cesarsky D. & Ryter C., 1998b, A&A 339, 194 Boulanger F., Beichman C., D´esert F.X. et al., 1988, ApJ 332, 328 Braine J., Combes F., Casoli F. et al., 1993, A&AS 97, 887 (CO ref.: Br) Brooks K.J., Burton M.G., Rathborne J.M., Ashley M.C.B. & Storey J.W.V., 2000, MNRAS 319, 95 Buta R. & Block D.L., 2001, ApJ 550, 243 Cayatte V., Kotanyi C., Balkowski C. & van Gorkom J.H., 1994, AJ 107, 1003 Cayatte V., van Gorkom J.H., Balkowski C. & Kotanyi C., 1990, AJ 100, 604 Cesarsky D., Lequeux J., Abergel A. et al., 1996a, A&A 315, L309; 1996b, A&A 315, L305 Chamaraux P., Balkowski C. & Fontanelli P., 1986, A&A 165, 15 (HI ref.: C2) Chamaraux P., Balkowski C. & G´erard E., 1980, A&A 83, 38 (HI ref.: C) Claussen M.J. & Sahai R., 1992, AJ 103, 1134 (CO ref.: CS) Cohen M., Allamandola L., Tielens A.G.G.M. et al., 1986, ApJ 302, 737 Combes F., Prugniel P., Rampazzo R. & Sulentic J.W., 1994, A&A 281, 725 (CO ref.: CP) Combes F., Dupraz C., Casoli F. & Pagani L., 1988, A&A 203, L9 (CO ref.: CD) Condon J.J., Frayer D.T. & Broderick J.J., 1991, AJ 101, 362 Contursi A., Lequeux J., Cesarsky D. et al., 2000, A&A 362, 310 Corwin H.G., de Vaucouleurs A. & de Vaucouleurs G., 1985, Southern Galaxy Catalogue Dale D.A., Silbermann N.A., Helou G. et al., 2000, AJ 120, 583 Dale D.A., Helou G., Silbermann N.A. et al., 1999, AJ 118, 2055 D´esert F.-X., Boulanger F. & Puget J.L., 1990, A&A 237, 215 Devereux N., 1987, ApJ 323, 91 Diaz A.I., Pagel B.E.J. & Wilson I.R.G., 1985, MNRAS 212, 737 Downes D., 1989, in "Evolution of Galaxies. Astronomical Observations", Les Houches Astrophysics School I, 351 Duley W.W. & Williams D.A., 1981, MNRAS 196, 269 Elfhag T., Booth R.S., Hoglund B., Johansson L.E.B. & Sandqvist A., 1996, A&AS 115, 439 (CO ref.: E) Eskridge P.B., Frogel J.A., Pogge R.W. et al., 2000, AJ 119, 536 Forster-Schreiber N.M., Laurent O., Sauvage M. et al., 2001, submitted to A&A Friedli D. & Benz W., 1993, A&A 268, 65 Garcia A.M., 1993, A&AS 100, 47 Garc´ıa-Barreto J.A., Franco J., Carrillo R., Venegas S. & Escalante-Ram´ırez B., 1996, RevMexAA 32, 89 Garc´ıa-Barreto J.A., Downes D., Combes F. et al., 1991, A&A 252, 19 (CO ref.: GD) Genzel R., Lutz D., Sturm E. et al., 1998, ApJ 498, 579 G´erin M., Nakai N. & Combes F., 1988, A&A 203, 44 (CO ref.: GN) Giard M., Bernard J.P., Lacombe F., Normand P. & Rouan D., 1994, A&A 291, 239 Giovanardi C., Krumm N. & Salpeter E.E., 1983, AJ 88, 1719 (HI ref.: GK) de Grijp M.H.K., Miley G.K. & Lub J., 1987, A&AS 70, 95 18 H. Roussel et al.: Impact of bars on MIR dust emission of spirals Guiderdoni B. & Rocca-Volmerange B., 1985, A&A 151, Maoz D., Koratkar A., Shields J.C. et al., 1998, AJ 116, 108 (HI ref.: GR) 55 Handa T., Nakai N., Sofue Y., Hayashi M. & Fujimoto M., 1990, PASJ 42, 1 (CO ref.: HN) Hawarden T.G., Mountain C.M., Leggett S.K. & Puxley P.J., 1986, MNRAS 221, Short Com. 41 Martin P. & Friedli D., 1997, A&A 326, 449 Martin P., 1995, AJ 109, 2428 Martinet L. & Friedli D., 1997, A&A 323, 363 Mathewson D.S. & Ford V.L., 1996, ApJS 107, 97 (HI ref.: Haynes M.P., van Zee L., Hogg D.E., Roberts M.S. & Mad- MF) dalena R.J., 1998, AJ 115, 62 (HI ref.: HZ) Mauersberger R., Henkel C., Walsh W. & Schulz A., 1999, Haynes M.P. & Giovanelli R., 1986, ApJ 306, 466 (HI ref.: A&A 341, 256 (CO ref.: M) HG) Heckman T.M., 1980, A&A 88, 365 (Research Note) Helfer T.T. & Blitz L., 1993, ApJ 419, 86 (CO ref.: HB) Helou G., 1986, ApJ 311, L33 Helou G., Hoffman G.L. & Salpeter E.E., 1984, ApJS 55, McLeod K.K. & Rieke G.H., 1995, ApJ 441, 96 Mirabel I.F. & Sanders D.B., 1988, ApJ 335, 104 (HI ref.: MS) Moshir M., Copan G., Conrow T. et al., 1989, IRAS Faint Source Catalog 433 (HI ref.: HH) Ho L.C., Filippenko A.V. & Sargent W.L.W., 1997, ApJS Mundell C.G. & Shone D.L., 1999, MNRAS 304, 475 Papoular R., Conard J., Giuliano M., Kister J. & Mille 112, 315 (nuc. type ref.: H) G., 1989, A&A 217, 204 Horellou C., Casoli F. & Dupraz C., 1995, A&A 303, 361 Pence W.D. & Blackman C.P., 1984, MNRAS 210, 547 (HI ref.: HC) (HI ref.: P) Huang J.H., Gu Q.S., Su H.J. et al., 1996, A&A 313, 13 Huchtmeier W.K. & Seiradakis J.H., 1985, A&A 143, 216 Phillips M.M. & Malin D.F., 1982, MNRAS 199, 905 Regan M.W., Vogel S.N. & Teuben P.J., 1997, ApJ 482, (HI ref.: HS) L143 Huchtmeier W.K. & Bohnenstengel H.D., 1981, A&A 100, Regan M.W., Teuben P.J. & Vogel S.N., 1996, AJ 112, 72 (HI ref.: HB) 2549 (HI ref.: RT) Huntley J.M., Sanders R.H. & Roberts W.W., 1978, ApJ Reynaud D. & Downes D., 1998, A&A 337, 671 (CO ref.: 221, 521 RD) Isobe T. & Feigelson E.D., 1992, ApJS 79, 197 Jorsater S. & van Moorsel G.A., 1995, AJ 110, 2037 (HI Rice W., Lonsdale C.J., Soifer B.T. et al., 1988, ApJS 68, 91 ref.: J) Richter O.G. & Huchtmeier W.K., 1987, A&AS 68, 427 Kamphuis J.J., Sijbring D. & van Albada T.S., 1996, (HI ref.: RH) A&AS 116, 15 (HI ref.: Ka) Roche P.F., Aitken D.K. & Smith C.H., 1989, MNRAS Keel W.C., 1984, ApJ 282, 75 (nuc. type ref.: K2) Keel W.C., 1983, ApJS 52, 229 (nuc. type ref.: K) Kenney J.D. & Young J.S., 1988, ApJS 66, 261 (CO ref.: KY) 236, 485 Rogstad D.H., Shostak G.S. & Rots A.H., 1973, A&A 22, 111 (HI ref.: RS) Roth J., Mould J. & Staveley-Smith L., 1994, AJ 108, 851 Kennicutt R.C., 1998, ApJ 498, 541 Koopmann R.A. & Kenney J.D.P., 1998, ApJ 497, L75 Kotilainen J.K., Reunanen J., Laine S. & Ryder S.D., (HI ref.: RM) Roussel H., Sauvage M., Vigroux L. & Bosma A., 2001a, accepted for A&A (Paper II) 2000, A&A 353, 834 Roussel H., Vigroux L., Bosma A., Sauvage M. et al. Krugel E., Steppe H. & Chini R., 1990, A&A 229, 17 (CO 2001b, A&A 369, 473 (Atlas) ref.: KS) Ryder S.D., Buta R.J., Toledo H. et al., 1996, ApJ 460, Krumm N. & Salpeter E.E., 1980, AJ 85, 1312 (HI ref.: 665 (HI ref.: RB) Kr) Sage L.J. & Isbell D.W., 1991, A&A 247, 320 (CO ref.: Kutner M.L. & Ulich B.L., 1981, ApJ 250, 341 Laurent O., Mirabel I.F., Charmandaris V. et al., 2000, SI) Sakamoto K., Okumura S.K., Ishizuki S. & Scoville N.Z., A&A 359, 887 1999, ApJ 525, 691 L´eger A. & Puget J.L., 1984, A&A 137, L5 Lindblad P.O., Hjelm M., Hogbom J. et al., 1996, A&AS Sakata A., Wada S., Onaka T. & Tokunaga A.T., 1987, ApJ 320, L63 120, 403 Sancisi R., Allen R.J. & Sullivan W.T, 1979, A&A 78, 217 Maia M.A.G., Pastoriza M.G., Bica E. & Dottori H., 1994, (HI ref.: S) ApJS 93, 425 (HI ref.: MP) Sandage A. & Bedke J., 1994, The Carnegie atlas of galax- Maia M.A.G., da Costa L.N., Willmer C., Pellegrini P.S. ies & Rit´e C., 1987, AJ 93, 546 (nuc. type ref.: MC) Maiolino R., Ruiz M., Rieke G.H. & Papadopoulos P., Sanders D.B. & Mirabel I.F., 1996, ARA&A 34, 749 Sanders D.B., Scoville N.Z. & Soifer B.T., 1991, ApJ 370, 1997, ApJ 485, 552 (nuc. type ref.: M) 158 (CO ref.: SS) H. Roussel et al.: Impact of bars on MIR dust emission of spirals 19 Sandqvist A., Jorsater S. & Lindblad P.O., 1995, A&A 295, 585 (CO ref.: SJ) Seigar M.S. & James P.A., 1998, MNRAS 299, 672 Sellgren K., 1984, ApJ 277, 623 Sheth K., Regan M.W., Vogel S.N. & Teuben P.J., 2000, ApJ 532, 221 (CO ref.: SR) Shioya Y., Tosaki T., Ohyama Y. et al., 1998, PASJ 50, 317 (CO ref.: ST) Soifer B.T., Boehmer L., Neugebauer G. & Sanders D.B., 1989, AJ 98, 766 Spoon H.W.W., Koornneef J., Moorwood A.F.M., Lutz D. & Tielens A.G.G., 2000, A&A 357, 898 Strong A.W., Bloemen J.B.G.M., Dame T.M. et al., 1988, A&A 207, 1 Sturm E., Lutz D., Tran D. et al., 2000, A&A 358, 481 Taniguchi Y., Ohyama Y., Yamada T., Mouri H. & Yoshida M., 1996, ApJ 467, 215 Telesco C.M., Dressel L.L. & Wolstencroft R.D., 1993, ApJ 414, 120 Theureau G., Bottinelli L., Coudreau-Durand N. et al., 1998, A&AS 130, 333 (HI ref.: TB) Thuan T.X. & Sauvage M., 1992, A&AS 92, 749 Tran D., 1998, PhD thesis, University of Paris XI Tubbs A.D., 1982, ApJ 255, 458 Tully R.B., 1988, Nearby Galaxies Catalog, Cambridge University Press (HI ref.: T) Uchida K.I., Sellgren K., Werner M.W. & Houdashelt M.L., 2000, ApJ 530, 817 de Vaucouleurs G., de Vaucouleurs A., Corwin H.G. et al., 1991, Third Reference Cat. of Bright Galaxies (RC3) Veilleux S., Bland-Hawthorn J. & Cecil G., 1999, AJ 118, 2108 Veilleux S., Kim D.C., Sanders D.B., Mazzarella J.M. & Soifer B.T., 1995, ApJS 98, 171 (nuc. type ref.: VK) V´eron-Cetty M.P. & V´eron P., 1986, A&AS 66, 335 (nuc. type ref.: V) Vila-Vilar´o B., Taniguchi Y. & Nakai N., 1998, AJ 116, 1553 (CO ref.: V) Vorontsov-Velyaminov B.A. & Arkhipova V.P., 1968, Mor- phological Catalog of Galaxies (part IV) Wiklind T., Henkel C. & Sage L.J., 1993, A&A 271, 71 (CO ref.: W1) Wiklind T. & Henkel C., 1989, A&A 225, 1 (CO ref.: W2) Whitworth A., 1979, MNRAS 186, 59 Young J.S, Xie S., Tacconi L. et al., 1995, ApJS 98, 219 (CO ref.: Y) 20 H. Roussel et al.: Impact of bars on MIR dust emission of spirals Table 1. General properties of sample galaxies. Virgo members are also named from the VCC catalog. Other galaxies belong to the field or loose groups, unless otherwise noted. Distances are from the NGC catalog of Tully (1988), taking into account the Virgo infall and assuming h100 = 0.75; morphological types, asymptotic blue magnitudes mBT and major diameters D25 at the isophote µB = 25 mag arcsec−2 are from the RC3 (de Vaucouleurs et al. 1991). name RA DEC (2000) D morph. mBT (Mpc) type D25 HI def.a (′) -- (b)b MH2 (log M⊙) -- (′′) nuclear type 8.05/7.65 (44) (E) 9.20/9.04 (44) (E) 8.91/8.76 (44) (GD) HII (A) L/HII (V) HII (V) L/HII (V) tidalc int. mbs am L/Sy (V/M) doa Sy1/2 (V/M) L/Sy2 (V/M) 5.13 2.88 5.50 2.40 9.33 11.22 6.46 4.57 6.61 3.16 7.08 2.82 11.22 11.22 12.88 3.16 28.84 19.95 3.39 -1.75 (P) -0.82 (T) 0.64 (MF) 0.98 (TB) 0.13 (MF) 0.22 (J) 0.87 (RB) -0.10 (RT) -1.24 (MP) -0.62 (RH) 0.88 (GR) 0.24 (HS) 0.77 (HS) -0.01 (AF) -2.40 (HB) 0.36 (S) 0.58 (HB) -0.18 (MF) -0.44 (T) 9.33 (17) (GN) 9.69/9.68 (44) (SJ) 8.36/8.24 (44) (BW) (5.3) (RD) 9.02/8.95 (44) (BW) 9.65 8.73/8.72 (21) (*) 9.14 (21) (W1) 7.27/7.24 (16) (ST) 8.95/8.99 (55) (HB) (16) (HN) 9.53/9.39 (55) (SR) 7.97/7.82 (55) (SI) 8.24 9.37/9.26 (44) (CS) 8.41/8.37 (16) (V) 8.12/7.89 (45) (KY) 8.58/8.43 (45) (KY) 8.66/8.41 (45) (KY) 0.33 (GR) 0.52 (GR) 1.23 (GR) 0.41 (GR) 1.84 (GR) 1.79 (GR) 1.48 (GR) 1.02 (HH) 2.53 (GR) 1.78 (GR) 5.13 9.77 5.62 2.00 5.62 3.63 2.34 2.29 8.99/8.88 (21) (CD) 8.51 8.37/8.05 (44) (Bo) 5.25 1.70 >1.73 (GK) 7.73/7.35 (33) (Bo) 2.95 1.68 (GR) 1.62 >1.60 (HG) 2.95 <1.04 (C) 4.57 9.55 5.89 2.09 2.14 2.57 2.88 4.90 4.27 2.11 (GR) 1.94 (GR) 1.92 (HH) 0.40 (HH) 1.05 (HH) 1.82 (GR) 0.16 (GR) 1.44 (GR) 8.32/7.97 (45) (KY) HII (H) HII (K) 8.56/8.28 (33) (E) L/HII (H) 8.98/8.77 (33) (Bo) 8.54/8.37 (33) (E) L/Sy2 (H) 8.11/7.79 (45) (KY) HII (K) HII (V) 8.32/8.12 (21) (Br) HII (K) 00 52 42−31 12.4 N289 00 59 50−07 34.7 N337 01 34 18−29 25.0 N613 02 38 33−06 40.7 N1022 02 46 19−30 16.4 N1097 03 33 37−36 08.3 N1365 † 03 42 01−47 13.3 N1433 ‡ 04 23 29+75 17.8 N1530 04 45 42−59 15.0 N1672 ‡ N4027 11 59 30−19 16.1 N4535 (V1555) 12 34 20+08 11.9 12 48 14−03 20.0 N4691 12 50 54+41 07.2 N4736 (M94) N5194 (M51) 13 29 53+47 11.8 N5236 (M83) 13 37 00−29 52.1 N5383 (Mrk281) 13 57 05+41 50.7 14 03 13+54 20.9 N5457 (M101) N6744 19 09 45−63 51.4 23 16 11−42 35.0 N7552 § Virgo cluster sample: 12 12 46+10 51.8 N4178 (V66) 12 13 48+14 53.7 N4192 (V92) 12 21 13+18 23.0 N4293 (V460) N4351 (V692) 12 24 02+12 12.4 N4388 (V836) # 12 25 47+12 39.7 12 25 56+18 12.9 N4394 (V857) N4413 (V912) 12 26 32+12 36.6 N4430 (V1002) 12 27 27+06 15.8 N4438 (V1043) 12 27 46+13 00.6 N4450 (V1110) 12 28 29+17 05.1 N4491 (V1326) 12 30 57+11 29.0 N4498 (V1379) 12 31 40+16 51.2 N4506 (V1419) 12 32 11+13 25.3 N4567 (V1673) 12 36 33+11 15.5 N4568 (V1676) 12 36 35+11 14.3 N4569 (V1690) 12 36 50+13 09.8 N4579 (V1727) 12 37 44+11 49.2 N4580 (V1730) 12 37 48+05 22.2 N4633 (V1929) 12 42 37+14 21.4 N4634 (V1932) 12 42 40+14 17.8 N4647 (V1972) 12 43 32+11 34.9 N4654 (V1987) 12 43 57+13 07.6 N4689 (V2058) 12 47 46+13 45.8 19.4 20.7 17.5 18.5 14.5 16.9 11.6 36.6 14.5 25.6 16.8 22.5 4.3 7.7 4.7 37.8 5.4 10.4 19.5 16.8 16.8 17.0 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 16.8 11.72 SBbc 12.06 SBd 10.73 SBbc 12.09 SBa 10.23 SBb 10.32 SBb 10.70 SBab 12.25 SBb 10.28 SBb SBdm 11.66 SABc 10.59 SB0/a 11.66 8.99 SAab SAbc 8.96 8.20 SABc 12.05 SBb 8.31 SABcd SABbc 9.14 11.25 SBab SBdm 11.90 SABab 10.95 SB0/a 11.26 SBab 13.03 SBab/ 11.76 11.73 SBb 12.25 SBab SBb 12.79 SA0/a 11.02 SAab 10.90 13.50 SBa 12.79 SABd 13.63 SBa SAbc 12.06 SAbc 11.68 SABab 10.26 10.48 SABb SABa 11.83 SABdm 13.75 SBcd/ 13.16 SABc 11.94 SABcd 11.10 SAbc 11.60 L/HII (V) HII (V) HII (V) HII (K) L (H) Sy2 (H) HII (V) HII (A) HII (H) L (V) HII (V) HII (H) L/HII (H) L (K) Sy2 (K) L (K) L (K) L (K) mbs am doa doa note doa pc doa H. Roussel et al.: Impact of bars on MIR dust emission of spirals 21 Table 1. continued. The last column indicates the name of the observation project: S for Sf glx or HI Q gal for IC 1953 (PI: G. Helou), and I for Irgal (PI: T. Onaka). name RA DEC (2000) D morph. mBT (Mpc) type D25 HI def.a (′) -- (b)b MH2 (log M⊙) -- (′′) nuclear type tidalc int. Supplementary galaxies: 02 33 34−39 02.6 N986 03 23 56−36 27.9 N1326 † 03 37 28−24 30.1 N1385 11 14 11+48 19.1 N3583 N3620 d 11 16 04−76 12.9 11 27 32+56 52.6 N3683 11 30 07+09 16.6 N3705 11 46 47−27 55.4 N3885 12 02 12+62 08.2 N4041 12 06 23+52 42.7 N4102 ♮ 12 33 30+08 39.3 N4519 (V1508) N4713 12 49 58+05 18.7 N5430 (Mrk799) d 14 00 46+59 19.7 N5786 d 14 58 57−42 00.8 N5937 d 15 30 46−02 49.8 15 36 32+16 36.5 N5962 N6156 d 16 34 52−60 37.1 N6753 d 19 11 23−57 02.9 N6824 d 19 43 41+56 06.6 20 34 52+60 09.2 N6946 22 10 12−16 39.6 N7218 N7418 22 56 36−37 01.8 N7771 (Mrk9006) d 23 51 25+20 06.7 03 33 42−21 28.8 IC1953 10 08 50−67 01.8 IC2554 23 28 43−41 20.0 IC5325 ESO317-G023 d 10 24 43−39 18.4 23.2 16.9 17.5 34.0 23.7 28.4 17.0 27.8 22.7 17.0 16.8 17.9 40.4 39.7 36.8 31.8 42.9 42.4 44.5 5.5 22.0 17.8 57.4 22.1 16.7 18.1 32.8 SBab 11.64 SB0+ 11.37 11.45 SBcd SBb 11.90 SBab SBc 13.15 SABab 11.86 SA0/a 11.89 SAbc 11.88 SABb 11.99 SBd 12.34 SABd 12.19 SBb 12.72 SBbc 12.17 SABb 13.11 SAc 11.98 SABc 12.30 SAb 11.97 SAb 13.00 SABcd 9.61 SBcd 12.70 SABcd 11.65 SBa 13.08 12.24 SBd SBbc 12.51 SABbc 11.83 SBa 13.93 3.89 3.89 3.39 2.82 2.75 1.86 4.90 2.40 2.69 3.02 3.16 2.69 2.19 2.34 1.86 2.95 1.58 2.45 1.70 11.48 2.51 3.55 2.51 2.75 3.09 2.75 1.91 0.47 (RM) -0.18 (HC) -0.14 (AH) -0.31 (T) 0.60 (RM) -1.46 (Ka) -0.28 (HZ) -1.08 (C2) -0.82 (T) 0.53 (T) -0.32 (GR) -0.82 (HZ) 0.81 (TB) -0.96 (RM) 0.60 (TB) 0.57 (Kr) L/HII (H) HII (H) HII (H) L/HII (H) HII (V) L (MC) HII (H) 9.52/9.43 (21) (KS) HII (K2) 8.86/8.80 (33) (E) 9.13/9.08 (21) (M) 7.63/7.29 (45) (Y) 9.48/9.39 (44) (E) S 8.43/8.33 (44) (W2) S 8.31/7.90 (44) (AC) HII (V) mbs S 9.45/9.09 (55) (SS) S 9.61/9.44 (44) (E) S S S S I S S S I S I S I S I S S S S S doa I S S L/HII (A) HII (V) 0.80 (Ka) -1.22 (RS) -0.70 (T) 0.28 (T) 9.86/9.58 (45) (Y) 0.28 (MS) 1.34 (T) 8.46/8.20 (44) (CP) -0.35 (RC3) 8.65/8.34 (44) (AB) 0.79 (T) HII (H) HII (VK) a The HI deficiency according to the definition and reference values of Guiderdoni & Rocca (1985). It is normalized by the dispersion in the field sample. Diameters are taken from the RC2 for consistency, and HI fluxes from the indicated references. NGC 4567/68 are unresolved in HI. When DHI/Dopt ≤ 1. and Def ≥ 0.5 in Cayatte et al. (1994), it corresponds to def > 1.2 here. b The given range represents the effect of varying the scale length of the CO distribution from once to twice that of infrared circumnuclear regions (see text). b is the beam HPBW of the observations used. (*) NGC 4535 was observed by A. Bosma, D. Reynaud and H. Roussel at the IRAM 30 m telescope. c Signs of tidal interaction. "doa": asymmetrical distortion of outer arms. "mbs": magellanic barred spiral. "am": amorphous. "pc": past collision. "note": On DSS images, the brightness peak is displaced by ≈ 10′′ ENE from the center of the regular outer isophotes and the NW outer disk seems depressed in stars and gas. † members of the Fornax cluster of galaxies. ‡ members of the Dorado group of galaxies. § member of the Grus quartet with NGC 7582/90/99. ♮ member of the spiral-rich Ursa Major cluster. # classified SA in the RC3, here considered a SB after the morphological arguments of Phillips & Malin (1982) and McLeod & Rieke (1995), and the recent kinematic analysis of Veilleux et al. (1999). d The distances of ESO 317-G023, NGC 6753 and 6156 were assumed to be those of the galaxy groups LGG 199, LGG 426 and LGG 407 (Garcia 1993); those of NGC 5786 and 7771 were estimated from the HI redshift, that of NGC 3620 from the CO redshift and those of NGC 5430, 5937 and 6824 from the optical redshift (with h100 = 0.75). 22 H. Roussel et al.: Impact of bars on MIR dust emission of spirals Table 2. Photometric results at 15 and 7 µm, obtained as described in the Atlas (total fluxes, diameter aperture used for central regions, fluxes inside this aperture and background levels). We warn the reader that the uncertainties can only be taken as order-of-magnitude values (see the Atlas), especially for galaxies of the third subsample belonging to the Sf glx project, with a very low number of exposures per sky position. Galaxies with no reported central fluxes have no identifiable central concentration: the radial surface brightness profile is consistent with a disk alone at our angular resolution (NGC 4580 rather shows a smooth central plateau and NGC 4634 is seen edge-on). For NGC 7552, we used only the maps with a 3′′ pixel size, because those at 6′′ are strongly saturated in both filters; for the other galaxies mapped with both pixel sizes, we used the 6′′ sampling because of the higher signal to noise ratio and the more reasonable field of view. name F15 tot F7 tot (mJy)a D CNR (′′) (kpc) F15 CNR F7 CNR (mJy)a b15 b7 (µJy arcsec−2) 68.5± 4.2 31.8± 2.2 557.1± 47.8 748.9± 81.2 1730.3± 92.6 3163.2± 420.2 117.2± 3.3 267.6± 10.2 1179.8± 74.7 31.7± 6.4 153.4± 15.8 730.5± 81.4 566.2± 39.2 2032.7± 33.4 3473.9± 200.2 185.8± 20.1 129.0± 9.2 26.6± 5.6 37.5± 1.8 28.5± 1.1 353.4± 31.3 350.8± 39.2 1285.1± 65.2 2019.0± 301.2 105.3± 2.4 217.0± 9.8 956.7± 69.8 32.4± 4.3 111.8± 14.2 510.9± 53.5 540.6± 49.0 1869.3± 40.5 2656.4± 203.3 171.3± 20.9 114.5± 4.7 50.3± 2.1 2292.1± 153.3 1251.8± 137.4 588.± 6. 679.± 6. 534.± 6. 744.± 9. 416.± 6. 370.± 4. 352.± 5. 345.± 5. 340.± 5. 900.± 4. 1043.± 5. 1374.± 5. 473.± 10. 412.± 10. 1096.± 5. 394.± 3. 361.± 2. 485.± 6. 565.± 7. N289 N337 N613 N1022 N1097 N1365 N1433 N1530 N1672 N4027 N4535 N4691 N4736 (−) N5194 N5236 # N5383 N5457 N6744 (−) N7552 # 327.8 ± 25.6 297.9 ± 24.0 1566.5 ± 104.0 802.3 ± 86.4 2269.2 ± 167.4 4436.7 ± 764.5 355.3 ± 41.0 606.1 ± 39.2 342.9± 14.7 336.1± 17.9 1473.3± 71.4 444.4± 45.3 2128.6± 125.4 3691.9± 616.6 381.3± 33.8 573.9± 39.1 2020.5 ± 123.0 1985.0± 129.2 676.7 ± 95.5 1127.9 ± 181.4 795.9 ± 185.6 4204.5 ± 240.6 8003.2 ± 493.5 20098.4 ± 803.7 332.6 ± 61.9 5424.3 ± 322.0 1497.4 ± 125.7 2767.6 ± 193.7 775.8± 68.2 1136.6± 68.9 613.5± 83.1 3913.9± 225.8 8598.7± 552.1 18474.9± 899.7 350.2± 62.1 6034.0± 116.7 2419.4± 52.3 1826.2± 168.5 Virgo cluster sample: N4178 N4192 N4293 (+) N4351 N4388 # N4394 N4413 N4430 N4438 (+) N4450 (+) N4491 N4498 N4506 N4567 † N4568 † N4569 N4579 N4580 N4633 N4634 N4647 N4654 N4689 181.5 ± 48.0 630.0 ± 99.6 188.6 ± 42.8 45.6 ± 26.3 1008.2 ± 244.0 139.0 ± 41.0 93.0 ± 31.4 98.0 ± 23.5 209.1 ± 34.7 169.7 ± 42.5 81.1 ± 25.2 94.6 ± 19.2 12.7 ± 5.2 293.4 ± 15.5 1099.0 ± 127.6 939.3 ± 125.1 619.2 ± 85.1 103.9 ± 24.2 30.0 ± 9.5 258.2 ± 40.7 472.3 ± 32.0 1018.6 ± 78.4 329.7 ± 37.4 228.5± 24.6 900.8± 68.3 159.5± 25.3 52.6± 8.7 499.4± 77.8 161.2± 19.1 89.3± 11.0 132.5± 13.9 231.9± 26.9 185.1± 14.6 30.5± 7.6 112.9± 11.8 21.1± 9.8 317.9± 16.4 1074.7± 64.8 843.5± 54.1 672.5± 37.5 102.6± 7.7 30.3± 9.1 278.3± 35.0 474.3± 17.2 1049.4± 42.9 340.9± 16.3 12.9 11.3 19.1 15.0 45.6 42.6 31.1 27.7 32.4 10.5 23.2 44.9 21.6 88.9 36.8 32.1 35.2 32.4 21.3 23.8 29.3 13.9 16.8 20.8 18.6 14.2 21.0 17.3 10.4 14.7 13.0 18.0 17.5 21.3 26.4 1.21 1.13 1.62 1.34 3.21 3.49 1.75 4.92 2.28 1.30 1.89 4.90 0.45 3.32 0.84 5.89 0.92 1.64 2.01 1.94 2.39 1.15 1.37 1.70 1.52 1.16 1.71 1.41 0.85 1.20 1.06 1.47 1.42 1.73 2.15 34.6± 4.7 132.3± 16.4 128.0± 34.9 14.3± 4.7 763.3± 244.9 19.7± 2.9 27.0± 4.0 33.4± 2.1 120.0± 23.8 50.2± 12.4 14.8± 1.0 224.7± 66.6 21.6± 3.2 22.4± 2.6 123.8± 22.7 103.3± 10.0 28.5± 4.5 73.4± 21.4 11.6± 2.2 7.5± 3.4 50.0± 5.7 230.2± 47.5 289.2± 88.1 152.2± 36.4 28.1± 2.9 18.3± 4.0 11.2± 1.0 7.6± 0.9 41.3± 2.3 172.2± 23.0 144.1± 23.4 96.4± 9.7 1051.± 3. 638.± 3. 916.± 3. 998.± 3. 994.± 3. 905.± 3. 992.± 3. 986.± 4. 906.± 3. 879.± 3. 1052.± 3. 876.± 3. 895.± 3. 979.± 3. 979.± 3. 871.± 3. 973.± 3. 1006.± 3. 830.± 3. 830.± 3. 849.± 3. 823.± 3. 796.± 3. 16.9 15.5 1.38 1.26 61.1± 6.9 92.5± 25.9 52.8± 2.7 82.9± 11.2 115.± 4. 99.± 4. 101.± 4. 139.± 4. 74.± 4. 72.± 2. 61.± 4. 61.± 4. 66.± 4. 171.± 2. 170.± 2. 237.± 2. 97.± 3. 74.± 12. 233.± 3. 68.± 2. 62.± 1. 103.± 3. 94.± 4. 178.± 2. 96.± 2. 170.± 2. 161.± 3. 159.± 2. 165.± 2. 158.± 3. 166.± 3. 178.± 2. 161.± 2. 169.± 3. 160.± 2. 149.± 3. 174.± 2. 174.± 2. 144.± 2. 170.± 2. 155.± 3. 139.± 2. 139.± 2. 183.± 2. 174.± 2. 132.± 2. H. Roussel et al.: Impact of bars on MIR dust emission of spirals 23 Table 2. continued. name F15 tot F7 tot (mJy)a D CNR (′′) (kpc) F15 CNR F7 CNR (mJy)a b15 b7 (µJy arcsec−2) Supplementary galaxies: 1050.0 ± 79.0 287.8 ± 58.8 782.3 ± 62.7 448.4 ± 52.5 1199.8 ± 408.6 755.5 ± 71.3 307.4 ± 73.6 396.0 ± 16.0 751.9 ± 128.1 1712.9 ± 561.0 N986 N1326 N1385 N3583 N3620 # N3683 N3705 N3885 N4041 N4102 # N4519 N4713 N5430 N5786 N5937 N5962 N6156 N6753 N6824 N6946 # 10651.6 ± 1767.2 N7218 N7418 N7771 I1953 I2554 I5325 ESO317 233.8 ± 67.6 209.4 ± 55.7 530.1 ± 136.1 380.2 ± 81.6 616.1 ± 128.9 508.8 ± 37.2 827.0 ± 149.2 646.8 ± 60.1 394.6 ± 98.1 273.5 ± 54.1 455.5 ± 113.3 615.2 ± 116.6 222.9 ± 37.0 887.6 ± 345.2 373.9 ± 42.2 287.2 ± 59.2 801.5± 12.1 284.9± 18.8 815.7± 27.1 425.6± 19.1 723.1± 89.2 791.1± 57.4 348.2± 32.6 342.9± 10.2 792.8± 164.5 808.8± 125.9 177.8± 18.3 223.6± 13.3 364.8± 153.7 345.8± 25.5 562.3± 94.6 485.5± 17.0 697.9± 118.0 586.4± 16.9 408.3± 101.3 11648.8± 678.6 260.6± 12.8 469.8± 32.9 526.8± 57.4 186.2± 13.7 733.5± 242.6 364.6± 17.4 239.8± 38.4 22.9 33.1 12.9 14.0 30.2 14.1 30.7 36.4 24.3 11.2 20.7 19.3 7.4 32.3 9.8 30.2 17.9 19.4 14.3 17.9 2.57 2.71 1.10 2.31 3.47 1.16 4.14 4.01 2.00 0.91 4.05 3.72 1.33 6.64 2.11 0.81 1.55 5.41 1.53 1.45 657.0± 67.1 258.4± 35.6 79.9± 7.2 87.2± 4.2 316.1± 11.5 212.3± 16.1 68.5± 4.6 70.0± 5.0 1087.2± 408.3 537.2± 74.7 28.6± 1.8 363.9± 8.9 482.8± 69.5 1419.2± 570.7 28.9± 2.1 293.0± 9.8 476.7± 123.7 462.3± 113.7 85.3± 24.3 20.3± 0.8 273.0± 80.3 120.1± 15.5 157.7± 90.1 152.5± 106.9 85.7± 5.4 85.2± 29.1 247.8± 7.3 81.5± 30.3 218.7± 3.1 74.4± 23.6 1730.4± 747.0 970.6± 277.5 38.6± 3.1 323.8± 76.1 116.2± 23.1 503.9± 369.7 31.1± 1.2 247.1± 33.9 34.4± 2.8 264.0± 183.4 19.9 3.17 222.2± 32.4 162.1± 20.7 372.± 4. 353.± 4. 385.± 4. 517.± 5. 386.± 4. 452.± 5. 686.± 6. 738.± 5. 428.± 6. 443.± 5. 1050.± 6. 845.± 6. 375.± 5. 823.± 7. 1140.± 8. 499.± 5. 457.± 6. 466.± 5. 335.± 5. 424.± 5. 1013.± 7. 640.± 5. 869.± 6. 429.± 3. 376.± 5. 537.± 5. 519.± 5. 63.± 2. 62.± 2. 73.± 2. 97.± 2. 81.± 2. 88.± 2. 96.± 2. 150.± 2. 96.± 4. 81.± 2. 183.± 2. 131.± 2. 81.± 4. 173.± 3. 234.± 4. 97.± 2. 86.± 5. 78.± 2. 69.± 4. 98.± 2. 162.± 2. 112.± 2. 180.± 3. 85.± 2. 81.± 4. 96.± 2. 112.± 3. a The conversion from flux densities to fluxes is: F (W m−2) = 10−14Fλ(Jy) × ∆ν(λ)(THz), with the filter widths ∆ν(15) = 6.75 THz and ∆ν (7) = 16.18 THz. # : nucleus saturated (for NGC 5236: both at 15 and 7 µm, but more severely at 15 µm since the same gain and integration time were used for both filters and since F15/F7 is above 1 in electronic units; for NGC 7552: slightly at 15 µm, but not at 7 µm; for the central pixel of NGC 4388: at 15 µm, but not at 7 µm; for NGC 3620: at 7 µm but not at 15 µm, which is possible because the integration time was respectively 5 s and 2 s; for NGC 4102 and 6946: both at 7 and 15 µm, but more severely at 7 µm, with the same configuration as NGC 3620). Thus, F15/F7 colors in central regions are respectively lower limits for NGC 5236, 7552 and 4388 and upper limits for NGC 3620, 4102 and 6946. (−) : The field of view is too small to allow a precise determination of the backgroud level and total fluxes are lower limits. The error bars are only formal. The comparison of our measurements with those of Rice et al. (1988) at 12 µm, inside the IRAS band 8 -- 15 µm which overlaps with our 5 -- 8.5 µm and 12 -- 18 µm bands, indicates that we miss of the order of 15% of total fluxes for NGC 4736 and between 15 and 45% for NGC 6744, provided IRAS fluxes are not overestimated as this is often the case for co-added observations. (+) : From their spectral energy distributions shown by Boselli et al. (1998), these galaxies probably have a non-negligible contribution from the Rayleigh-Jeans tail of cold stars to their 7 µm emission. We did not attempt to remove this contribution, because it would require a careful modelling of stellar populations. † : The disks of these galaxies slightly overlap in projection. We attempted to separate them by the means of a mask defined visually, but the disk fluxes are much more uncertain than estimated.
astro-ph/0112444
1
0112
2001-12-19T11:59:33
A sub-mm imaging survey of ultracompact HII regions
[ "astro-ph" ]
We present the preliminary results of a sub-mm imaging survey of ultracompact HII regions, conducted with the SCUBA bolometer array on JCMT.
astro-ph
astro-ph
**TITLE** ASP Conference Series, Vol. **VOLUME**, **PUBLICATION YEAR** **EDITORS** A sub-mm imaging survey of ultracompact HII regions Mark Thompson Centre for Astrophysics & Planetary Science, School of Physical Sciences, University of Kent, Canterbury, UK Jenny Hatchell Max Planck Institut fur Radioastronomie, Bonn, Germany Geoff Macdonald Centre for Astrophysics & Planetary Science, Department of Electronics, University of Kent, Canterbury, Kent, UK Tom Millar Astrophysics Group, Department of Physics, UMIST, Manchester, UK Abstract. We present the preliminary results of a sub-mm imaging sur- vey of ultracompact HII regions, conducted with the SCUBA bolometer array on the JCMT. 1. Introduction Ultracompact (UC) HII regions are currently the best known tracer of massive YSOs and represent the earliest confirmed stage of massive star formation. In excess of 150 UC HII regions have been detected, mainly by radio surveys. Whilst the environments of UC HII regions are known very well on the small scale (a few arcseconds) they are not well known on scales over 40′′. This is because most UC HIIs have, to date, been observed using either interferometers (to gain information on small scales at the expense of large scales) or by single- position large-beam (typically 40′′ or worse) spectroscopy. To redress this issue we recently undertook an imaging survey of over 100 UC HII regions using SCUBA on the JCMT, which enables us to rapidly map (with high resolution) the dust emission from the clumps in which the UC HIIs are embedded. 2. The survey SCUBA is mainly comprised of two bolometer arrays which (almost) instanta- neously sample a 2′ field simultaneously at 450 and 850 µm. The instrument is very sensitive, with the results that we were able to image each UC HII down to a 1σ noise level of typically 50 mJy at 850 µm in only 3 minutes. Our source sample was drawn from the UC HII region catalogues of Wood & Churchwell 1 2 Thompson, Hatchell, Macdonald & Millar G10.84−2.59 G23.86+0.15 n o i t a n i l c e D 46:30 −20:47:00 30 48:00 30 49:00 53:30 −7:54:00 30 55:00 30 56:00 n o i t a n i l c e D 16 14 12 18:19:10 08 06 18:34:30 28 26 24 22 20 Right ascension Right ascension Figure 1. of the UC HII regions. Holes in the image are removed noisy pixels. 850 µm images from the survey. Crosses mark the locations (1989) and Kurtz, Churchwell & Wood (1994), comprising some 140 UC HII re- gions in all. Our motivations for the survey were to i) investigate the large-scale structure of the dust clumps embedding the UC HIIs; ii) fill the existing gap in the spectral energy distribution at sub-mm wavelengths; iii) search for other unknown dust clumps in the field of view, possibly harbouring massive YSOs or protostars in different evolutionary states and iv) identify hot molecular cores via their strongly peaked sub-mm emission (Hatchell et al. 2000). 3. Preliminary results In total we observed 106 out of the 140 UC HII regions in the catalogues. Sub- mm emission was detected in 80% of the sample. The morphology of the dust clumps is surprisingly linear: over half of the singly-peaked clumps are elongated along one axis (e.g. G10.84 in Fig. 1) and over half of multiply-peaked clumps take the form of cores string along a ridge (e.g. G23.96 in Fig. 1). We also identified a large number of previously unknown dust clumps in the field of view, many of which are not associated with radio continuum or embedded IR sources and may contain massive protostars in an earlier stage to that of the UC HII (see Gibb et al., this proceedings for details of BIMA follow-up observations). In addition we identified 15 UC HIIs associated with strongly peaked sub-mm continuum which may contain hot molecular cores. References Hatchell, J., Fuller G.A., Millar, T.J., Thompson, M.A., Macdonald, G.H. 2000, A&A, 357, 637 Kurtz, S., Churchwell, E., Wood, D.O.S 1994, ApJS, 91, 659 Wood, D.O.S., Churchwell, E. 1989, ApJS, 69, 831
astro-ph/0103039
1
0103
2001-03-02T13:51:42
The Ursa Major Cluster of Galaxies. III. Optical observations of dwarf galaxies and the luminosity function down to M_R=-11
[ "astro-ph" ]
Results are presented of a deep optical survey of the Ursa Major Cluster, a spiral-rich cluster of galaxies at a distance of 18.6 Mpc which contains about 30% of the light but only 5% of the mass of the nearby Virgo Cluster. Fields around known cluster members and a pattern of blind fields along the major and minor axes of the cluster were studied with mosaic CCD cameras on the Canada-France-Hawaii Telescope. The dynamical crossing time for the Ursa Major Cluster is only slightly less than a Hubble time. Most galaxies in the local Universe exist in similar moderate density environments. The Ursa Major Cluster is therefore a good place to study the statistical properties of dwarf galaxies since this structure is at an evolutionary stage representative of typical environments yet has enough galaxies that reasonable counting statistics can be accumulated. The main observational results of our survey are: (i) The galaxy luminosity function is flat, with a logarithmic slope alpha = -1.1 for -17 < M_R < -11, from a power-law fit. The error in alpha is likely to be less than 0.2 and is dominated by systematic errors, primarily associatedd with uncertainties in assigning membership to specific galaxies. This faint end slope is quite different to what was seen in the Virgo Cluster where alpha=-2.26. (ii) Dwarf galaxies are as frequently found to be blue dwarf irregulars as red dwarf spheroidals in the blind cluster fields. The density of red dwarfs is significantly higher in the fields around luminous members than in the blind fields. The most important result is the failure to detect many dwarfs. If the steep luminosity function claimed for the Virgo Cluster were valid for Ursa Major then in our blind fields we should have found about 1000 galaxies with -17 < M_R <-11 where we have found two dozen.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 27 October 2018 (MN LATEX style file v1.4) The Ursa Major Cluster of Galaxies. III. Optical observations of dwarf galaxies and the luminosity function down to MR = −11 Neil Trentham1, R. Brent Tully2 and Marc A. W. Verheijen3 1 Institute of Astronomy, Madingley Road, Cambridge, CB3 0HA. 2 Institute for Astronomy, University of Hawaii, 2680 Woodlawn Drive, Honolulu HI 96822, U. S. A. 3 NRAO-Array Operations Center, P. O. Box 0, Socorro NM 87801, U. S. A. 27 October 2018 ABSTRACT Results are presented of a deep optical survey of the Ursa Major Cluster, a spiral-rich cluster of galaxies at a distance of 18.6 Mpc which contains about 30% of the light but only 5% of the mass of the nearby Virgo Cluster. Fields around known cluster members and a pattern of blind fields along the major and minor axes of the cluster were studied with mosaic CCD cameras on the Canada-France-Hawaii Telescope. The dynamical crossing time for the Ursa Major Cluster is only slightly less than a Hubble time. Most galaxies in the local Universe exist in similar moderate density environments. The Ursa Major Cluster is therefore a good place to study the statistical properties of dwarf galaxies since this structure is at an evolutionary stage representative of typical environments yet has enough galaxies that reasonable counting statistics can be accumulated. The main observational results of our survey are: (i) The galaxy luminosity function is flat, with a logarithmic slope α = −1.1 for −17 < MR < −11 from a power-law fit. The error in α is likely to be less than 0.2 and is dominated by systematic errors, primarily associated with uncertainties in assigning membership to specific galaxies. This faint end slope is quite different to what was seen in the Virgo Cluster, where α = −2.26 ± 0.14. (ii) Dwarf galaxies are as frequently found to be blue dwarf irregulars as red dwarf spheroidals in the blind cluster fields. The density of red dwarfs is significantly higher in the fields around luminous members than in the blind fields. The most important result is the failure to detect many dwarfs. If the steep lumi- nosity function claimed for the Virgo Cluster were valid for Ursa Major then in our blind fields we should have found ∼ 103 galaxies with −17 < MR < −11 where we have found two dozen. There is a clear deficiency of dwarfs compared with the ex- pectations of hierarchical clustering theory. It is speculated that the critical difference between the Virgo and Ursa Major clusters is the very different dynamical collapse times, which probably straddle the timescale for reionization of the Universe. Dwarf galaxies in the proto-Virgo environment probably formed before the epoch of reioniza- tion. The equivalent dwarf halos in the proto-Ursa Major environment probably only formed after the epoch of reionization, when the conditions for star formation were inhospitable. Key words: galaxies: luminosity function -- galaxies: photometry -- galaxies; clusters: individual: Ursa Major 1 0 0 2 r a M 2 1 v 9 3 0 3 0 1 0 / h p - o r t s a : v i X r a 1 INTRODUCTION The galaxy luminosity function φ(L), defined as the num- ber density of galaxies per unit luminosity L, is an important probe of the physical processes that leads to galaxy forma- c(cid:13) 0000 RAS tion. The luminosity function depends on both the primor- dial fluctuation spectrum and on the physics governing star formation. The luminosity function has been a popular di- agnostic since it is so straightforward to measure, at least at high luminosities, and it is expected to tell us something 2 Neil Trentham, R. Brent Tully and Marc A. W. Verheijen about a more fundamental parameter, the mass function. Early theoretical attempts were successful at reproducing the general form of the luminosity function (e.g. White & Rees 1978), which decreases monotonically with increasing luminosity, and decreases very steeply indeed above some characteristic luminosity L∗ ∼ 2 × 1010 h−2 75 LB ⊙ (Schechter 1976; where h75 = H◦/(75 km s−1 Mpc−1); H◦ is the Hubble Constant). More recent theoretical models (e.g. Baugh et al. 1998, Somerville & Primack 1999, Diaferio et al. 1999) have been based on a semi-analytic approach to the study of galaxy formation and are able to make more detailed pre- dictions. The luminosity function of galaxies in the field is nor- mally determined based on an imaging survey over a chosen angular region of the sky with follow-up redshift measure- ments of the galaxies (e.g. the Las Campanas Redshift Sur- vey of Lin et al. 1996 and the Autofib survey of Ellis et al. 1996). From Hubble's law, one has distances and hence can compute luminosities and construct a luminosity func- tion for a cone volume with the observer at the origin. The luminosity functions measured this way tend to be very well- determined at the bright end but poorly-determined at the faint end. Most galaxies in a magnitude-limited sample are distant luminous galaxies, not nearby low-luminosity galax- ies. The clumpiness of the distribution of galaxies and the fact that the faintest galaxies are drawn from a very small region means the normalization of the faint end relative to the brighter end is usually dubious. Imaging to extremely deep limits does not help since the angular coverage on the sky is then necessarily small -- for example there are only two galaxies in the faintest bin of the Keck survey of Cowie et al. (1996). Galaxies in the Local Group are known down to very faint limits (MV ∼ −8.5), but the local luminos- ity function (van den Bergh 1992, 2000) suffers from poor counting statistics at all luminosities, in addition to possi- ble incompleteness at the very faint end. For example, the Cetus dwarf (Whiting et al. 1999), with MV ∼ −10, was discovered only last year. An alternative approach to measuring the luminos- ity function down to very low luminosities is as follows. The vast majority of low-luminosity galaxies (the "dwarf galaxies") have low surface-brightnesses and follow (al- beit with some scatter) the absolute magnitude vs. central surface-brightness correlation shown in Figure 1 of Binggeli (1994). Red dwarf spheroidal (dSph) and blue dwarf irregu- lar (dIrr) galaxies both have azimuthally-averaged light pro- files that are exponential, and follow the same absolute mag- nitude vs. central surface-brightness correlation (Binggeli & Cameron 1991, Binggeli 1994). Therefore if one finds a low surface-brightness galaxy of a given apparent magnitude in a cluster, it is far more likely to be a low-luminosity member of that cluster than a high-luminosity background galaxy. The converse is true if one finds a high surface-brightness galaxy of the same apparent magnitude. These statements can be made more quantitative by observations of blank sky fields -- there is a marked absence of low surface-brightness galaxies in these fields, as we shall see later in this paper. One place where this kind of technique has been employed very suc- cessfully is in the Virgo Cluster, where Phillipps et al. (1998) find a very steep galaxy luminosity function, with α ∼ −2 (here α is the logarithmic slope of the luminosity function: φ(L) ∼ Lα) between MR = −14 and MR = −11, close to the predicted slope of the galaxy mass function from Press & Schechter (1974) theory, assuming the cold dark matter fluctuation spectrum of Bardeen et al. (1986). These results probe significantly deeper than the well-known Virgo lumi- nosity function of Sandage, Binggeli & Tammann (1985; see also Impey, Bothun & Malin 1987), or the studies of nearby groups of Tully (1988) or of Ferguson & Sandage (1991), all of whom found far shallower luminosity functions. The Virgo results, however, do not necessarily tell us anything about the field luminosity function, which is what is important for cosmology. That is because the Virgo Cluster is a dense envi- ronment where the galaxies formed early and where galaxy- galaxy interactions probably played an anomalously impor- tant role in shaping present-day galaxy properties. So we now apply these techniques to the Ursa Major Cluster, a large but diffuse cluster of spiral galaxies at dis- tance of 18.6 Mpc (Tully & Pierce 2000), close to the Virgo Cluster, and attempt to determine the luminosity function of low-luminosity galaxies there. This cluster is very loosely held together and has no appreciable X-ray halo. It is quite different from clusters of elliptical galaxies like Virgo (itself not a particularly rich cluster). The velocity dispersion of the Ursa Major Cluster is low (148 km s−1, compared to 715 km s−1 for Virgo), and its mass is about 1/20 of the mass of the Virgo Cluster (Tully 1987b). The results for the Ursa Major Cluster are expected to be be far more represen- tative of what the field galaxy luminosity function will look like at the faint end because most galaxies in the Universe exist in diffuse spiral-rich environments. A deep optical survey of sufficient angular area for good Poisson statistics is made possible in nearby clusters like Ursa Major by the advent of mosaic CCDs on large tele- scopes. In this work we used the UH8K (Metzger, Luppino & Miyazaki 1995) and CFH12K (Cuillandre et al. 1999) mo- saic cameras on the Canada-France-Hawaii Telescope (here- after CFHT) to image the cluster along its major and minor axes, along with a number of pointed observations around known cluster members. Additionally we have obtained HI images with the Very Large Array (hereafter VLA) of the fields along the major and minor axes. In this paper we present all the optical data. The radio data are presented elsewhere (Verheijen et al. 2001). The fields around known cluster members already have HI data available, from the Westerbork Synthesis Radio Telescope (hereafter WSRT; Verheijen 1998). We also have taken pointed multicolour images of probable new members and present those data here. Additionally, we will be able to use the results to de- termine the colour distribution of low-luminosity galaxies in the cluster, which constrains star formation histories, and give attention to the morphology-density relation of dwarf galaxies in the Ursa Major Cluster. It should be clear from our data whether the dwarf galaxies congregate around the giant galaxies or are more uniformly distributed within the cluster, and whether or not the answer depends on the dwarf galaxy morphology, HI mass, and colour. In the local Uni- verse, these three properties certainly play an important role: red gas-poor dwarf spheroidals cluster around giant galaxies, whereas blue gas-rich dwarf irregulars are less cor- related in position with giant galaxies (Binggeli, Tarenghi & Sandage 1990). This paper is organized as follows. In Section 2 we de- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The Ursa Major Cluster of Galaxies 3 scribe the observations and basic data processing. In Section 3 we describe how we perform photometry and outline how we identify plausible cluster candidate members. In Section 4 we look at the sample so constructed in more detail and describe how confident we are about each identification. Us- ing this sample, we determine luminosity functions, colour distributions, and morphology-density relations in Sections 5 -- 7. Finally in Section 8 we summarize and attempt to put all the results together to obtain a coherent picture of the Ursa Major Cluster. 0 2 OBSERVATIONS AND DATA REDUCTION 2.1 General strategy Our basic observing strategy for this project was as follows: in March 1996 and March 1999 we observed regions in the Ursa Major Cluster using large-format mosaic CCDs with the intention of finding low luminosity galaxies, which we identified based on their low surface brightnesses. A number of background fields were taken for comparison purposes. Pointed three filter observations of the candidates that we identified were then taken in February 2000 using a single- chip CCD so that colours could be measured. In 1996 we mostly observed fields centered on known cluster members (see Figure 1 and Table 1). In 1999 we observed contiguous fields along the major and minor axes of the cluster (again see Figure 1 and Table 1; the field designations we used are presented there). All the fields studied in 1999 were also observed with the HI line receiver at the VLA and all but the 'blank' fields observed in 1996 were observed with the HI line receiver at the WSRT. A number of dwarf candidates turned out to be HI gas-rich and we are therefore able to confirm that they really are members from their observed velocities. For the regions of the sky we study, Galactic extinction is small: E(B − V ) < 0.05 mag but varies slightly from field to field and we use the measurements of Schlegel, Finkbeiner & Davis 1998 to correct our data for this effect. 2.2 First Observing Run (1996) Images were taken of the fields listed in Table 1. All images were taken at the prime focus of the CFHT on Mauna Kea, using the UH8K mosaic camera, a mosaic of eight 4K × 2K CCDs (Metzger et al. 1995; scale 0.22 arcsec pix−1; total field of view 0.5 degree × 0.5 degree). The total area sur- veyed was then 2.2 square degrees. Each field was imaged as a set of three 1200 seconds exposures in the R-band, dithered by up to an arcminute to reject cosmic rays and bad pix- els, and to ensure that regions that fell in the gaps between the CCDs on any particular exposure were imaged in at least one other exposure. The R-band filter was chosen for this survey to maximize the magnitude limit for the detec- tion of low surface-brightness galaxies with this instrumen- tal setup; the unthinned CCDs had low quantum efficiency at shorter wavelengths and airglow emission contaminates at longer wavelengths. All images were dark-subtracted and flat-fielded using twilight sky flats. Due to geometric distor- tions arising from the large size of the camera, we did not combine the images until after the galaxy detection stage c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 2000 0 (RA offset)/s -2000 Figure 1. The field positions, shown as offsets from the cluster center which we take to be (α(2000), δ(2000)) = (11h59m28.3s, 49◦05′18′′). The large squares represent the UH8K fields observed in 1996 (see Table 1). The letters/numbers repre- sent the CFH12K fields observed in 1999; the designations here are as in Table 1. The size of the CFH12K field is shown in the rectangle at the bottom left. The small open circles represent known NGC members. The small open squares represent known UGC members. The small open triangles represent other known members. The large circle represents the approximate extent of the Ursa Major Cluster; i.e. the region within 7.5 degrees of the cluster center as defined above. The positions of the nearby 12−3 and 14−4 Groups (Tully 1987a) are also shown. since these geometric transforms alter the noise statistics in a complex way (see Section 3; to compensate for this effect we ran the detection programs at a low significance threshold so as not to miss any marginally-detected galax- ies). Instrumental magnitudes were computed from observa- tions of standard stars, and the photometry was converted to the Cousins R magnitude system of Landolt (1992). Images taken under non-photometric conditions were calibrated ini- tially using the data of Tully et al. (1996, hereafter Paper I) and eventually using the data we obtained during the third observing run of the current program (see Section 2.4). The median seeing was 1.0 arcseconds. 2.3 Second Observing Run (1999) Images were taken of the fields listed in Table 1. All im- ages were taken at the prime focus of the CFHT on Mauna Kea, using the CFH12K mosaic camera, a mosaic of twelve 4K × 2K CCDs (Cuillandre et al. 1999; scale 0.22 arcsec pix−1; total field of view 0.7 degree × 0.5 degree). The to- tal area surveyed was then 15.8 square degrees. Each field was imaged for 420 seconds. Subsequent exposures were pro- gressively shifted a half -- field diameter. Hence, most parts of the sky along the major and minor axes were imaged twice. The projection of camera gaps or flaws shift between ex- 4 Neil Trentham, R. Brent Tully and Marc A. W. Verheijen Table 1. Fields observed Field α (2000) δ (2000) Known Members (R mag)∗ Other galaxies (R mag)† 1996 Blank 1 Blank 2 NGC 3953 field UGC 6917 field NGC 3998 field 11 37 32.4 11 52 38.9 11 53 48.6 11 56 30.4 11 56 47.6 56 55 14 51 09 03 52 23 36 50 23 16 55 23 12 UGC 6930 field NGC 3992 field 11 57 19.1 11 57 36.2 49 16 55 53 14 29 NGC 4100 field UGC 7176 field Blank 3 12 06 08.1 12 10 55.3 12 24 52.3 49 34 59 50 15 49 50 43 23 1999 A01 11 46 09.2 56 09 07 NGC 3953 (9.66) UGC 6917 (12.16) NGC 3998 (9.55) NGC 3972 (11.90) NGC 3990 (12.08) UGC 6930 (11.71) NGC 3992 (9.55) UGC 6923 (12.97) UGC 6969 (14.32) UGC 6940 (15.65) NGC 4100 (10.62) UGC 7176 (15.61) A02 A03 A04 A05 A06 A07 A08 A09 A10 A11 A12 A13 A14 A15 A16 A17 A18 A19 A20 A21 A22 A23 A24 A25 A26 A27 A28 A29 A30 A31 A32 A33 11 47 05.4 11 48 00.2 11 48 53.9 11 49 46.3 11 50 37.6 11 51 27.8 11 52 17.0 11 53 05.2 11 53 52.3 11 54 38.6 11 55 23.9 11 56 08.3 11 56 51.9 11 57 34.6 11 58 16.6 11 58 57.8 11 59 38.3 12 00 18.0 12 00 57.0 12 01 35.4 12 02 13.1 12 02 50.2 12 03 26.8 12 04 02.7 12 04 37.9 12 05 12.8 12 05 47.0 12 06 20.7 12 06 54.0 55 42 15 55 15 20 54 48 25 54 21 28 53 54 30 53 27 30 53 00 29 52 33 26 52 06 23 51 39 19 51 12 13 50 45 06 50 17 59 49 50 50 49 23 40 48 56 30 48 29 18 48 02 06 47 34 52 47 07 38 46 40 23 46 13 08 45 45 51 45 18 34 44 51 16 44 23 58 43 56 39 43 29 19 43 01 58 12 07 26.7 12 07 58.9 12 08 30.7 42 34 37 42 07 16 41 39 54 NGC 3953 (9.66) UGC 6840 (13.35) UGC 6922 (13.65) UGC 6956 (13.83)∗5 UGC 6917 (12.16) UGC 6930 (11.71) NGC 4051 (9.88) NGC 4111 (9.95) NGC 4117 (12.47) NGC 4118 (14.82) UGC 7094 (13.70) UGC 7089 (12.77) 1203+43 (15.79) NGC 4143 (10.55) NGC 3977 (13.4)∗0 NGC 3888 (11.8)∗1 NGC 3898 (B = 11.6)∗2 NGC 3850 (13.2)∗3 UGC 6828 (13.5)∗4 PC1200+4755 (16.0)∗6 CGCG215−022 (12.8)∗7 UGC 7069 (14.4)∗8 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The Ursa Major Cluster of Galaxies 5 Field α (2000) δ (2000) Known Members (R mag)∗ Other galaxies (R mag)† B04 B05 B06 B07 B08 B09 B10 B11 B12 B13 B14 B15 B16 11 37 49.8 11 41 07.6 11 44 26.8 11 47 47.2 11 51 08.8 11 54 31.8 11 57 55.8 12 03 04.0 12 06 30.9 12 09 58.7 12 13 27.6 12 16 57.4 12 20 28.1 47 27 29 47 38 00 47 48 08 47 57 55 48 07 22 48 16 25 48 25 06 48 37 25 48 45 08 48 52 29 48 59 26 49 06 00 49 12 09 1148+48 (16.12) NGC 3985 (12.26) NGC 3811 (12.1)∗9 NGC 4047 (11.9)∗10 UGC 7358 (13.2)∗11 ∗ Magnitudes from Paper I. More details about the known members can be found there. † Magnitudes from this work. When the galaxy was not completely in our field of view, magnitudes are taken from the NASA Extragalactic Database, and the reader is referred there for the original sources. When R magnitudes are not available, B magnitudes are quoted. ∗0 NGC 3977 is a background grand-design spiral galaxy with a heliocentric velocity of 5722 km s−1. ∗1 NGC 3888 is a background late-type galaxy at z = 0.008. ∗2 NGC 3898 is centered just off this field to the East, and is only partly visible. It is a member of the nearby 12-3 Group and has a heliocentric velocity of 1176 km s−1. ∗3 NGC 3850 is a member of the nearby 12-3 group, with a heliocentric velocity of 1140 km s−1 (Verheijen et al. 2000). ∗4 UGC 6828 is a grand-design spiral, presumably background. ∗5 UGC 6956 is centered just off this field to the East, and is only partly visible in our field of view. ∗6 PC1200+4755 is a foreground emission-line galaxy with a spectroscopic redshift z = 0.002 (Schneider et al. 1994). ∗7 CGCG215−022 is a giant red early-type galaxy, presumably background. ∗8 UGC 7069 is a luminous flat late-type galaxy not seen in HI (Verheijen et al. 2000) and therefore presumably not a cluster member. ∗9 NGC 3811 is a background late-type galaxy at z = 0.010. ∗10 NGC 4047 is a background late-type galaxy at z = 0.011. ∗11 UGC 7358 is a background late-type galaxy at z = 0.012. posures so all parts of the sky along these axes were im- aged at least once. The fields are designated by the letter A along the major axis (along with a number between 1 and 33 in decreasing order of declination) and by the let- ter B along the minor (along with a number between 4 and 16 in increasing order of right ascension; fields B01 -- 03 and B17 -- 22 were covered in the VLA survey but not the opti- cal one). Exposures were taken in the R-band in order to maintain consistency with the 1996 data. The areas covered by the combined datasets are shown in Figure 1. All images were bias-subtracted (the dark current was negligible) and flat-fielded using twilight sky flats. Instrumental magnitudes were computed from observations of standard stars, and the photometry was again converted to the Cousins R magni- tude system of Landolt (1992). Some of the images were taken under marginally non-photometric conditions. Origi- nally, photometric zero-points were obtained using the data on luminous galaxies of Paper I with interpolation across the overlapping fields. However, the CFHT12K mosaic cam- era is large enough that extinction due to cirrus could vary across the detector by up to 0.2 magnitudes along the long axis of the chip (this happened in one or two of the images, where the average extinction was as much as one magnitude; in most images it was much smaller). The entire dataset was recalibrated once the data from the 2000 observing run was reduced. The median seeing was 0.8 arcseconds. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 2.4 Third Observing Run (2000) Images were taken of the candidate cluster members that we identified in the 1996 and 1999 data (see Sections 3 and 4). All images were taken at the f/8 Cassegrain focus of the University of Hawaii 2.2 m Telescope, also on Mauna Kea, using a Tektronix 2048 × 2048 thinned CCD (scale 0.22 arc- sec pix−1; field of view 7.5 arcmin × 7.5 arcmin). All data were taken under photometric conditions. Each object was imaged for 360 seconds with a B filter, 180 seconds with an R filter, and 180 seconds with an I filter. All images were bias-subtracted (the dark current was negligible) and flat-fielded using twilight sky flats. Instrumental magnitudes were computed from observations of a large number (about 100 in each filter) of standard stars, and the photometry was again converted to the Johnson (B) -- Cousins (RI) magni- tude system of Landolt (1992). The photometric zero-points were accurate to approximately one percent. The median seeing was 0.8 arcseconds. 3 3 PHOTOMETRY AND MEMBERSHIP CONSIDERATIONS In a diffuse environment like the Ursa Major Cluster, at R > 16 the number of background galaxies is higher than the number of cluster members (both the 1996 and 1999 datasets). Statistical background subtraction, as done in dis- 6 Neil Trentham, R. Brent Tully and Marc A. W. Verheijen tant clusters (e.g. Trentham 1998), will not be sufficient to allow us to correct for background contamination. We need to take the morphologies of the galaxies that we detect into account. As outlined in Section 1, members can be identi- fied based on the magnitude vs. central surface-brightness relation of Binggeli (1994). For galaxies of a given appar- ent magnitude, cluster members that are dwarfs have lower surface-brightnesses and larger sizes than background gi- ants of the same apparent magnitude. The light is much less concentrated in the dwarfs. Our approach is to make a detailed study of the 1996 dataset plus a number of back- ground fields (two blank UH8K fields near the cluster -- the "Blank 1" and "Blank 3" fields -- taken in 1996 and nine CFH12K fields taken in 1999) and derive a condition for membership candidature based on the concordance be- tween measured light concentrations (quantified according to Binggeli's correlation) and the existence of a number of objects (19 in the 1996 dataset) having lower light concen- trations than any galaxies of equivalent apparent magnitude in any of the background fields. We then apply these criteria to the far bigger 1999 dataset. Note that dwarf spheroidals and dwarf irregulars follow the same magnitude vs. central surface-brightness correlation (Binggeli 1994) so that if the objects in the 1996 fields are mostly dwarf spheroidals but those in the 1999 fields are mostly dwarf irregulars, this will not bias our results. Our application to the 1996 data was then as fol- lows. Objects were detected above local sky (Poisson noise- dominated, 1σ between 27 and 28 R mag arcsec−2) at a low (2σ) significance level using the FOCAS detection algo- rithm (Jarvis & Tyson 1981; Valdes 1982, 1989). For each detected galaxy we then define an inner concentration pa- rameter based on aperture R magnitudes: ICP = R(< 4.4 arcsec) − R(< 2.2 arcsec), and an outer concentration parameter: OCP = R(< 12 arcsec) − R(< 6 arcsec). Both the ICP and OCP are more negative for galaxies of larger scale length, which for a given apparent magnitude equates to galaxies of lower surface-brightness. Both are close to zero for stars, since the seeing was always much less than 2.2 arcsec (the seeing was always good enough that its effect on the concentration parameters for all the galaxies that we consider here was negligible). These concentration parameters characterize the light distribution on physical scales between about 0.2 kpc and 1 kpc (in Ursa Major one arcsecond is equivalent to 0.09 kpc). The concentration pa- rameters for galaxies in the 1996 dataset are presented in Figure 2. We now define two conditions: and ICP < −1.1 OCP < −0.4. (C1) (C2) Only one object in the background fields satisfies these con- ditions (which we later excluded as being an object that would mimic a candidate member had it turned up in a cluster field, under point (v) in the list below) yet a signifi- cant number of low surface-brightness objects in the cluster fields do satisfy these conditions. One would expect normal dwarf galaxies to satisfy both conditions, given Figure 1 of Binggeli (1994 -- see the lines in Fig. 2). We therefore regard objects that satisfy both conditions as possible cluster mem- bers. Figure 2 shows that the differential between members and non-members as defined by these conditions is not com- pletely clean: some objects which satisfy these conditions are certainly background objects due to their being either (i) grand-design luminous spirals with very negative concen- tration parameters due to star formation in spiral arms at large distance from the galaxy center, (ii) merging galaxies with no well-defined center, (iii) objects with a neighboring galaxy or star that did not get separated into two objects by the detection algorithm, (iv) extremely flat edge-on galaxies (dwarfs are relatively stubby), or (v) low surface-brightness material that seems to be debris or ejecta associated with a nearby giant galaxy. These cases are easily excluded from the sample. A few objects are less straightforward to ex- clude as background objects since they may show some of these signatures at a low level; these are the objects that we categorize "2" or "3" later in this section and discuss individually in Section 4. In the entire 1996 dataset there were 4154 extended objects with a 6 arcsecond aperture magnitude R(6) < 21.5, only 130 of which were left after imposing condition C1. Those that pass condition C1 were mostly normal background galaxies towards the lower end of the surface-brightness distribution (condition C1 was a very conservative one), but also included clusters members, ob- jects satisfying (i) -- (v) above, and some borderline cases. All these objects were studied by eye and the different kinds of objects identified. After imposing condition C2, there were 40 objects left, 17 of which were excluded as being candidate members based on criteria (i) through (v). The remaining 23 were identified as probable or possible members. Four were marginal as regards satisfying these conditions and/or have some morphological hints that they could be background ob- jects. We were less certain about these and categorize them "2" or "3" later in this section. In addition there was one object amongst the 130 that satisfied C1 but failed to sat- isfy C2 that is probably background, but in our judgment could be an extreme cluster member; we categorize this one "3" as discussed later in this section. Finally, we excluded one very flat low surface-brightness object that appeared to move with respect to the background galaxies between ex- posures; this may have been a small comet or alternatively an internal reflection within the camera. The prescription outlined in the previous paragraph can now be applied to the 1999 dataset, which is much larger (there were about 30,000 galaxies with R(6) < 21.5). All objects satisfying C1 and all that were within 0.1 magni- tudes of satisfying both C1 and C2 were looked at indi- vidually. Contaminant objects according to the criteria (i) through (v) above were excluded upon inspection by eye. Additionally we found a number of bright late-type galaxies that satisfied both C1 and C2 but were not seen in HI, and questioned their membership on these grounds (see Section 4 for further details). Most of the difficult cases are found at the north end of the major axis, close to the 12-3 Group (Tully 1987a), and could arise from contamination by that group. Note that the HI observations did not cover the entire velocity range of the 12 -- 3 group. It will be clear from the previous two paragraphs that we are somewhat more confident about the membership pos- sibilities of some objects than others. We therefore introduce c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Table 2. New members and candidates Galaxy Rating Comments† Field∗ α (2000) δ (2000) ICP OCP RTOT MR B − R R − I umd aperture arcsec The Ursa Major Cluster of Galaxies 7 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 2 3 2 1 3 3 0 3 0 1 1 1 1 2 2 1 3 2 1 2 0 1 0 3 1 3 1 3 2 1 1 1 1 1 1 1 1 0 1 0 1 2 0 0 3 2 0 1 1 1 1 1 1 2 2 1 1 2 2 1 AB A C C D E E A A E E E EF C AEG B H B E E E E E E I E E E C C I E E E I AB G B4 B5 B6 A2 B6 A1 A2 A4 B7 A4 A4 A3 A3 A3 A2 A1 A4 A3 A3 A4 A8 N3953, A10 B8 A11 A12 B9 N3998 U6917 B9 N3998 N3998 N3998 N3992 N3992 N3992 N3992 N3998 A17 N3998 A17 N3992 N3992 A16 A20 A19 A21 A25 A29 A27 A28 N4100 N4100 N4100 A30 N4100 A30 A28 B12 U7176 U7176 11 36 04.7 11 39 27.5 11 44 10.9 11 45 16.3 11 45 57.9 11 46 19.0 11 46 35.0 11 46 35.9 11 46 49.6 11 46 53.4 11 47 13.9 11 47 22.3 11 47 33.5 11 48 12.6 11 48 43.6 11 48 45.4 11 49 18.7 11 49 26.3 11 50 45.6 11 50 50.7 11 51 53.6 11 53 09.2 11 53 11.1 11 53 52.3 11 54 27.6 11 54 40.8 11 55 38.2 11 55 53.3 11 55 59.1 11 56 09.4 11 57 01.6 11 57 03.1 11 57 03.8 11 57 05.6 11 57 21.0 11 57 36.6 11 58 02.7 11 58 11.6 11 58 13.7 11 58 26.0 11 58 34.3 11 58 47.9 11 59 57.6 12 00 35.3 12 00 35.6 12 02 21.3 12 02 43.7 12 04 10.3 12 04 49.9 12 05 24.8 12 05 45.5 12 06 05.2 12 06 05.8 12 06 26.0 12 06 26.7 12 06 27.9 12 07 09.4 12 07 44.7 12 09 28.1 12 09 34.4 47 31 14 −1.21 −0.50 47 34 10 −1.18 −0.68 48 02 24 −1.00 −0.63 55 34 31 −1.64 −0.50 47 37 16 −1.13 −1.12 56 02 17 −1.07 −0.40 55 49 16 −1.24 −0.69 54 47 55 −1.24 −0.72 48 05 33 −1.14 −0.56 54 40 10 −1.77 −0.76 54 35 57 −1.25 −0.80 55 26 10 −1.36 −0.76 55 11 03 −1.21 −0.49 55 10 26 −1.35 −0.55 55 55 43 −1.28 −0.77 56 01 56 −1.23 −0.51 54 58 15 −0.94 −0.42 55 15 13 −1.40 −1.20 55 06 45 −1.20 −0.56 54 46 00 −1.05 −0.71 53 05 59 −1.20 −0.66 52 11 22 −1.45 −0.97 48 11 18 −1.23 −0.61 51 29 38 −1.14 −0.56 51 20 05 −1.26 −0.66 48 13 49 −1.11 −0.34 55 22 04 −1.57 −0.82 50 31 12 −1.23 −0.26 48 12 02 −1.03 −0.58 55 15 54 −1.46 −0.64 55 25 10 −1.23 −0.76 55 25 12 −1.39 −0.78 53 18 03 −1.47 −0.49 53 26 28 −1.16 −0.66 53 13 35 −1.42 −0.67 53 10 01 −1.34 −0.85 55 14 48 −1.16 −0.99 48 52 55 −1.41 −1.06 55 23 16 −1.25 −0.81 48 57 36 −1.36 −0.71 53 20 44 −1.17 −0.74 53 27 14 −1.31 −0.57 49 33 50 −1.36 −0.76 47 46 24 −1.48 −1.08 47 58 10 −1.00 −0.37 47 07 36 −1.13 −0.44 45 11 28 −1.16 −0.72 43 39 17 −1.15 −0.53 44 26 34 −1.28 −0.64 43 42 32 −1.46 −1.04 49 42 54 −1.30 −0.98 49 28 47 −1.14 −0.94 49 25 37 −1.32 −0.68 42 54 33 −1.13 −0.44 49 33 24 −1.39 −0.85 43 01 14 −1.13 −0.27 43 59 15 −1.14 −0.43 48 51 23 −1.36 −0.61 50 12 42 −1.50 −0.51 50 26 02 −1.46 −0.96 17.69 −13.66 17.77 −13.58 19.13 −12.22 18.84 −12.51 18.84 −12.51 16.67 −14.68 16.12 −15.23 18.53 −12.82 17.80 −13.55 18.50 −12.85 18.14 −13.21 17.65 −13.70 17.62 −13.73 18.07 −13.28 16.56 −14.79 19.15 −12.20 19.77 −11.58 17.29 −14.06 18.58 −12.77 19.96 −11.39 15.75 −15.60 15.27 −16.08 17.46 −13.89 16.67 −14.68 17.68 −13.67 19.45 −11.90 19.45 −11.90 18.90 −12.45 18.85 −12.50 18.13 −13.22 15.66 −15.69 15.75 −15.60 19.10 −12.25 17.72 −13.63 17.22 −14.13 18.71 −12.64 18.08 −13.27 14.76 −16.59 15.69 −15.66 17.68 −13.67 16.79 −14.56 20.58 −10.77 14.43 −16.92 13.70 −17.65 19.58 −11.77 17.40 −13.95 14.55 −16.80 19.35 −12.00 18.28 −13.07 15.17 −16.18 18.83 −12.52 19.76 −11.59 17.75 −13.60 18.67 −12.68 20.56 −10.79 19.48 −11.87 19.61 −11.74 19.80 −11.55 19.58 −11.77 17.19 −14.16 0.32 ± 0.08 1.00 ± 0.06 0.08 ± 0.12 0.80 ± 0.07 0.23 ± 0.21 0.84 ± 0.15 0.28 ± 0.41 1.18 ± 0.34 0.28 ± 0.22 0.70 ± 0.12 0.31 ± 0.03 1.05 ± 0.02 0.29 ± 0.03 0.84 ± 0.02 0.42 ± 0.22 1.43 ± 0.21 0.28 ± 0.03 0.91 ± 0.07 0.25 ± 0.22 1.22 ± 0.18 0.76 ± 0.28 1.22 ± 0.30 0.45 ± 0.23 1.31 ± 0.21 0.38 ± 0.07 1.26 ± 0.06 0.18 ± 0.19 0.43 ± 0.11 0.19 ± 0.08 1.09 ± 0.06 0.29 ± 0.26 1.26 ± 0.22 0.02 ± 0.33 0.60 ± 0.19 0.96 ± 0.27 0.98 ± 0.26 1.32 ± 0.16 −0.14 ± 0.24 1.59 ± 0.51 0.13 ± 0.55 0.31 ± 0.01 0.72 ± 0.01 0.41 ± 0.03 0.99 ± 0.02 0.44 ± 0.08 0.59 ± 0.05 0.20 ± 0.04 0.61 ± 0.02 1.01 ± 0.10 0.22 ± 0.14 1.05 ± 0.18 −0.40 ± 0.35 0.83 ± 0.55 0.13 ± 0.44 0.75 ± 0.08 0.16 ± 0.12 1.46 ± 0.19 −0.21 ± 0.28 1.41 ± 0.16 0.48 ± 0.16 0.37 ± 0.01 1.17 ± 0.01 0.39 ± 0.01 1.21 ± 0.01 0.00 ± 0.23 1.24 ± 0.16 0.26 ± 0.22 1.27 ± 0.19 1.05 ± 0.03 0.34 ± 0.04 0.15 ± 0.21 0.53 ± 0.13 0.41 ± 0.13 1.24 ± 0.11 0.29 ± 0.01 0.81 ± 0.01 0.48 ± 0.01 1.56 ± 0.01 0.58 ± 0.16 0.15 ± 0.25 0.44 ± 0.02 1.27 ± 0.02 0.35 ± 0.52 1.36 ± 0.49 0.40 ± 0.01 0.98 ± 0.01 0.38 ± 0.04 0.85 ± 0.03 1.04 ± 0.15 0.13 ± 0.21 0.36 ± 0.04 1.19 ± 0.04 1.25 ± 0.01 0.47 ± 0.01 1.17 ± 0.23 −0.32 ± 0.42 1.07 ± 0.17 −0.44 ± 0.34 1.18 ± 0.05 0.39 ± 0.05 1.13 ± 0.31 0.17 ± 0.41 1.28 ± 0.38 −1.82 ± 1.43 0.36 ± 0.07 1.27 ± 0.06 0.18 ± 0.24 1.15 ± 0.18 0.74 ± 0.45 0.09 ± 0.68 0.43 ± 0.31 1.06 ± 0.26 0.16 ± 0.41 1.21 ± 0.32 0.03 ± 0.34 0.87 ± 0.21 0.23 ± 0.21 0.86 ± 0.15 0.79 ± 0.08 0.20 ± 0.12 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 18 18 6 6 12 12 12 24 24 6 12 6 24 6 6 6 24 6 6 12 6 6 6 6 6 6 6 6 6 6 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 8 Neil Trentham, R. Brent Tully and Marc A. W. Verheijen Galaxy Rating Comments† Field∗ α (2000) δ (2000) ICP OCP RTOT MR B − R R − I aperture arcsec umd 61 62 63 64 65 J EK E 3 1 1 1 2 U7176 U7176 U7176 U7176 B14 12 09 47.8 12 11 13.3 12 11 22.7 12 11 34.9 12 14 43.7 50 03 51 −1.13 −0.78 50 06 41 −1.39 −0.87 50 16 11 −1.39 −0.92 50 26 13 −1.39 −0.89 49 01 28 −1.01 −0.41 19.34 −12.01 18.71 −12.64 15.11 −16.24 18.92 −12.43 20.28 −11.07 0.55 ± 0.15 −0.10 ± 0.29 0.09 ± 0.27 1.10 ± 0.19 1.23 ± 0.01 0.40 ± 0.01 1.37 ± 0.26 0.03 ± 0.34 0.94 ± 0.52 −0.18 ± 0.84 6 6 24 6 6 ∗ See Table 1 for the designations. For the 1996 data, the fields are named for luminous giant galaxies. For the 1999 data, the fields are labelled according to their position along the major or minor axis. † Legend for the third column: A: No HI detection, although clearly irregular/late-type morphology B: weak evidence for a bar C: high central surface-brightness (i.e. small physical size) if a cluster dwarf D: possibly associated with a nearby galaxy E: smooth morphology; probably a dSph F: huge extended halo G: blue colour H: extreme low surface-brightness I: irregular morphology; probably a dIrr J: extremely flat; probably background K: two nucleii; probably a recent merger the following subjective rating scheme, based on our own as- sessment, for all objects we find other than those presented in Paper I. Candidates are characterized "0" to "3", where "0": membership confirmed from HI data (Verheijen et al. 2000, Verheijen et al. 2001). There were no new HI de- tections in the 1996 fields, but nine new detections in the 1999 fields; "1": probable member, but no HI detection; "2": possibly a member, but conceivably background; "3": probably background, but conceivably a member. Our judgments are based primarily on by how clearly each object satisfies C1 and C2, the morphological criteria (i) through (v), and on the implications of no HI detection in bright galaxies with distinctly irregular morphologies. The justifications for our rating of each object are listed in the next section, along with comments on some objects that we excluded as being possible candidates. For each candidate we estimate its total R magnitude as follows. (i) We compute the flux within some isophote slightly above the extraction limit set by the sky back- ground. This isophote corresponded to the 1.8σ isophote in the 1996 data and the 1.5σ isophote in the (less deep) 1999 data. For a few extremely large low surface-brightness galaxies, the galaxies could reliably be followed beyond this isophote to a fainter one. (ii) We compute the amount of light that falls below the specified isophote at large radius by fitting an exponential light profile to a part of the galaxy where there are no condensations and extrapolating this light profile beyond the last fitted isophote to infinity (Pa- per I). The results are estimates of the total, as opposed to isophotal magnitudes. The uncertainties in these mag- nitudes depend on the details of the light distribution be- yond the isophotal radius for each galaxies, which are un- certain, but are likely to be far less than the bin size (2 mag) that we shall use in computing luminosity functions. We only consider galaxies whose isophotal magnitudes are brighter than R = 21.5. At fainter magnitudes it is difficult to rate with confidence any galaxies as "1" since low surface- brightness galaxies in the background fields begin to appear at these faint limits. The faintest magnitude for our lumi- nosity function is therefore not determined by detection con- straints, but by where we lose the ability to distinguish clus- ter members from background galaxies on surface-brightness grounds. Apparent magnitudes were then converted to ab- solute magnitudes assuming a distance of 18.6 Mpc, cor- responding to a distance modulus of 31.35. Colours (from the 2000 data) were computed from apertures centered on the R-band galaxy center. Using aperture magnitudes en- sures that we are probing the same stellar populations in all filters. The aperture sizes were between 6 and 24 arcsec- onds. Larger apertures were used for bigger galaxies so as to improve the signal-to-noise. Errors in the colours are dom- inated by Poisson sky noise and are large for the faintest galaxies. The apertures were always large enough so that differential seeing effects between the different filters were always negligible. Clearly our methods of identifying members a priori bias us towards selecting a particular kind of galaxy -- nor- mal dwarf galaxies. By far most local dwarfs are normal dwarf spheroidals or dwarf irregulars (Binggeli 1994), but it is still instructive to see how we fare with other kinds of low luminosity galaxies like blue compact dwarfs (BCDs). Most BCDs have irregular morphologies (Telles, Terlevich & Melnick 1997) and extended star-formation and/or low surface-brightnesses. We expect these objects to satisfy both C1 and C2 and so to appear in our samples. As a test, we placed the well-studied BCD galaxy UGC 6456 (Lynds et c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The Ursa Major Cluster of Galaxies 9 Figure 3. The CFHT R-band images for candidate members, along with their membership ratings in brackets. In all images north is up and east is to the left. The horizontal bar in each image represents 6 arcseconds. be extremely difficult to find anywhere; e.g. Disney 1976). If they did exist and were gas-rich we would find them in our HI survey. The disks of Malin 1 (Bothun et al. 1987), F568−6 (Bothun et al. 1990) and GP1444 (Davies, Phillipps & Disney 1988), would have easily turned up in our HI sur- vey, and probably in our optical survey as well. It is only gas-poor extreme low surface-brightness galaxies we need to worry about (the stars in such hypothetical objects could be gravitationally bound by a dark matter halo, for example). (ii) We also miss galaxies with smooth de Vaucouleurs light profiles but faint (> 20 mag arcsec−2) central surface- brightnesses, because they look like background ellipti- cals. Recall from the fundamental plane (see Kormendy & Djorgkvoski 1989) that higher luminosity ellipticals have fainter central surface-brightnesses, so that if we observe an elliptical galaxy with a moderate central surface-brightness, it is a priori likely to be a background luminous galaxy. Galaxies with the surface brightness profiles of elliptical galaxies but that lie so far away from the fundamental plane as to cause misidentifications with the background appear to be rare, but there was at least one in our survey. Markar- ian 1460, a Blue Compact Galaxy, failed to satisfy C1 and C2 very substantially, but is a known cluster member based on optical spectroscopy (Pustilnik et al. 1999) and on an HI detection (Verheijen et al. 2001). Only this one object is known in the cluster (Trentham, Tully & Verheijen 2001), although cases with less HI could be missed. 4 SAMPLE As described in the previous section, we now have a list of possible members, each with a rating "0" to "3" depending on our confidence regarding membership. These are listed in Table 2, along with the coordinates and photometric prop- erties. Images from the CFHT data are presented in Figure 3. The objects are numbered in order of increasing right as- cension. For the galaxies we detect we make various specific comments in the third column of Table 2; these often indi- cate why a particular galaxy received a particular rating. We do not attempt classifications as dwarf spheroidals or irregulars (although see the list below for more obvious cases) because such classifications are not reliable based on morphological information alone (eg, some ellipsoidal, tex- tureless dwarfs have HI and some lumpy dwarfs are not de- tected in HI). Figure 2. The concentration parameters for galaxies in the 1996 dataset, as defined in the text, as a function of apparent magni- tude. The top panel shows the inner concentration parameter for all extended objects with R(6) < 21.5 as a function of of ap- parent R magnitude measured in a 6 arcsecond circular aper- ture. The dots represent objects we regard as as background. The open symbols represent objects classified members or possi- ble members (circles = rated "1"; squares = rated "2"; triangles = rated "3"). The dashed line is the median predicted position for dwarf spheroidal galaxies, assuming the magnitude vs. surface- brightness correlation of Binggeli (1994), exponential light profiles (Binggeli & Cameron 1991), B − R = 1.5 (Trentham 1998 Section 5 and references therein), and the same observing conditions as for our 1996 data. The lower panel shows the outer concentration parameter for all extended objects as a function of of apparent R magnitude measured in a 12 arcsecond circular aperture. Only ob- jects in the upper panel with R(4.4)−R(2.2) < −1.1 are included. Objects believe to be background are labeled as filled circles or squares, depending on the absence or presence of a nearby object. The open symbols and the dashed line have the same meaning as in the upper panel. Background objects close to or below the dashed line were identified as such by morphology (e.g. grand- design spiral, or extreme flatness -- see points (i) through (v) in Section 3 of the text) or by the presence of a companion making the outer concentration parameter anomalously negative. al. 1998) in the Ursa Major Cluster and were able to recover it using the above strategy. In addition, we expect BCDs to be detected in HI and so turn up in the VLA sample. There are, however, two kinds of objects that we do miss, given our selection criteria. (i) We do not find extreme low surface-brightness disks, which have central surface-brightnesses below 27 R mag arcsec−2. No such galaxies are known (although they would c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 10 Neil Trentham, R. Brent Tully and Marc A. W. Verheijen Below are comments on some of the more interesting objects that satisfied the above selection criteria but we excluded from our sample. Where no names exist, we de- rive them from the J2000 coordinates: hhmm.m(+/-)ddmm. These are: NGC 3850 -- This galaxy easily satisfied both C1 and C2, but its heliocentric velocity of 1140 km s−1 (Verheijen et al. 2001) places it in the nearby 12-3 Group and not in Ursa Major. We therefore exclude it from the sample; 1145.9+5605 -- This very bright (R ∼ 14) late-type galaxy would surely have been seen in HI were it in the cluster, but it was not. It therefore must be a background object. Additionally we see signs of weak spiral structure, which would be consistent with this being a luminous background late-type galaxy; 1154.7+5053 -- This late-type peculiar galaxy (R ∼ 17.5) shows weak spiral structure, including a number of knots embedded in one arm, which are probably HII regions. This object was not seen in HI and so is presumably background; 1154.2+5053 -- This bright (R ∼ 16) late-type peculiar galaxy was not seen in HI and so is presumably background; PC1200+4755 -- This irregular low surface-brightness galaxy easily satisfied both C1 and C2, but it is a fore- ground object given its spectroscopic redshift z = 0.002 (Schneider, Schmidt & Gunn 1994); 1152.9+4754 -- This bright (R ∼ 17.5) late-type peculiar galaxy was not seen in HI and so is presumably background. It has two nucleii and weak spiral structure, which would support this interpretation; 1155.5+4846 -- This bright (R ∼ 15.5) interacting peculiar galaxy was not seen in HI and so is presumably background; 1207.1+4259 -- This bright (R ∼ 16) late-type peculiar galaxy was not seen in HI and so is presumably background. It has some spiral structure with many condensations (pre- sumably HII regions) embedded in the arms, which would support this interpretation; 1206.4+4226 -- This bright (R ∼ 16.5) flattened late-type galaxy was not seen in HI and so is presumably background; 1209.6+4201 -- This bright (R ∼ 14.5) late-type galaxy was not seen in HI and so is presumably background. There is weak evidence for spiral structure, which would support this interpretation; 1213.4+4901 -- This bright (R ∼ 16.5) late-type galaxy was not seen in HI and so is presumably background. Again there is weak evidence for spiral structure, which would sup- port this interpretation. Additional notes on these and other galaxies that were ex- cluded are presented in the notes to Table 1. 5 LUMINOSITY FUNCTIONS The luminosity functions are presented in Figure 4. It is immediately apparent that wherever we set the borderline between members and non-members (based on our ratings), the Ursa Major luminosity function is much shallower than the Virgo one. Values of α were computed from power-law fits to the data and the results are presented in Table 3. The value of α appropriate to Ursa Major is about −1.1, with an uncertainty of about 0.2. In comparison, in the Virgo Cluster α = −2.26 ± 0.14 from a similar power-law fit over UMa 1996 -- members from Tully et al. 1996 + galaxies ranked ''1'' UMa 1999 -- members from Tully et al. 1996 + galaxies ranked ''0'' to ''1'' Virgo (outer area sample from Phillipps et al. 1998) UMa 1996 -- members from Tully et al. 1996 + galaxies ranked ''1'' to ''3'' UMa 1999 -- members from Tully et al. 1996 + galaxies ranked ''0'' to ''3'' Virgo (outer area sample from Phillipps et al. 1998) Figure 4. The luminosity function of the Ursa Major Cluster. The 1996 data is offset horizontally by 0.2 mag to permit a clearer comparison with the 1999 data. The Virgo Cluster data by Phillipps et al. (1998) and the 1996 data are normalized to have the same number of MR = −15 galaxies rated "0" or "1" as in the 1999 data. The histograms represent the luminosity func- tions for the bright-galaxy sample of Papers I and II, assum- ing a completeness limit of MR = −18 and a normalization set by scaling the total number of galaxies with MR < −18 to be the same as for the 1999 dataset. The dashed lines represent the best-fitting power-laws to the 1999 data (the filled circles) for MR > −18; these have α = −0.95 (upper panel) and α = −1.16 (lower panel). The dotted-dashed lines represent the best-fitting Schechter (1976) functions to the joint Paper I + II (MR < −18) R, α∗) = and 1999 (MR > −18) datasets. These fits have (M ∗ (−21.44, −1.01) (upper panel) and (−21.71, −1.16) (lower panel). the range −16 < MR < −11.5 (Phillipps et al. 1998). Note, however, that in this study of the Virgo Cluster there was no culling of the sample for background objects, as in the present paper, which might lead to a slight overestimate of the Virgo numbers. The Virgo and Ursa Major luminosity functions are therefore highly inconsistent with each other. There is one caveat: the Ursa Major sample may be incomplete if there exist substantial numbers of either (see Section 3) gas-poor extreme low-surface-brightness galaxies or low-luminosity galaxies that look like distant elliptical or S0 galaxies due to the combination of central surface-brightnesses around 19 R mag arcsec−2 and smooth de Vaucouleurs light profiles (like Markarian 1460). In order for the Ursa Major luminosity function to be the same as the Virgo one, we would need to have missed many hundred such objects. Objects of the c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Table 3. Power law fits to lumiosity functions The Ursa Major Cluster of Galaxies 11 Galaxies rated "1" Galaxies rated "1" -- "2" Galaxies rated "1" -- "3" 1996 data −16.85 < MR < −10.85 −18.85 < MR < −10.85 α = −1.00 ± 0.18 α = −1.08 ± 0.16 α = −1.04 ± 0.17 α = −1.10 ± 0.15 α = −1.10 ± 0.16 α = −1.16 ± 0.14 1999 data −16.85 < MR < −10.85 −18.85 < MR < −10.85 α = −0.90 ± 0.18 α = −1.00 ± 0.12 α = −1.07 ± 0.14 α = −1.12 ± 0.10 α = −1.15 ± 0.13 α = −1.18 ± 0.09 Known members from Tully et al. (1996) are included in all the fits, in addition to the new candidates. first type do not appear to be common: on going to fainter magnitudes we did not find proportionately more galaxies with extremely low surface brightnesses. The lowest surface brightness object that we detected was umd 27, which had an average surface brightness of 24.6 R mag arcsec−2 within the 20% light radius, somewhat brighter than the 1-σ limit quoted in Section 3. Given that the only case of the second type known is Markarian 1460, we do not regard as a serious worry either. Since the Ursa Major Cluster is a diffuse unevolved clus- ter of spiral galaxies and since most galaxies in the Uni- verse reside in such environments, our results may be rep- resentative of the field luminosity function as well, down to MR = −11 (MB ∼ −10). The Local Group is another ex- ample of such an environment and has a similarly shallow luminosity function (α = −1.1; van den Bergh 1992), albeit with large errors due to poor counting statistics. More nega- tive values of α (i.e. steeper luminosity functions) have been found in the spectroscopic surveys of Lin et al. (1996) and Marzke et al. (1994), although these values of α come from a Schechter (1976) function fit to the bright galaxies, not a power-law fit to galaxies as faint as the ones we observe here. The values of α we present are only valid for the magnitude ranges identified in Table 3, which only overlap marginally at the bright end with the spectroscopic samples. Given this fact, and the difference in the way α is computed from the data, the two results are not inconsistent. The fraction of HI -- detectable objects is decreasing to- ward fainter absolute magnitudes. Nevertheless some of the faintest optical galaxies (like umd 40) do have significant amounts of cold gas. A more quantitative investigation of this phenomenon is presented in Verheijen et al. (2001). 6 COLOURS The colours of the new galaxies are listed in Table 2 and presented in Figures 5 and 6. In Figure 7 we present a his- togram of the B −R colours of galaxies whose errors in B −R are less than 0.2 mag. The galaxies we consider here all have −18 < MR < −10 if they lie in Ursa Major, hence qualify as dwarfs (see Binggeli 1994). Typical values are 1.2 < B − R < 1.6 for dSph galax- ies and B − R < 1.1 for dIrr galaxies (see Coleman, Wu & Weedman 1980 and Section 5 of Trentham 1998; see also c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Figure 5. Colour-colour diagram for the 1996 data, with colours computed as described in the text. The symbols have the follow- ing meanings: filled circles -- confirmed members (from Paper I); open circles -- galaxies rated "1"; open squares -- galaxies rated "2"; open triangles -- galaxies rated "3". Larger symbols repre- sent galaxies of higher luminosity. The lines are the stellar popu- lation evolutionary tracks from the models of Bruzual & Charlot (1993), computed assuming a metallicity of Z = 0.008 (0.4 solar) and a Salpeter (1995) initial mass function with mass limits of 0.1 and 100 M⊙. The different lines have the following meanings: solid line -- instantaneous burst; short dashed line -- continuous star formation model; long dashed line -- exponentially decaying star formation model with e-folding timescale of 1 Gyr. The open pentagons on the instantaneous burst model lines represent the colours at 0.5, 1, 2, 4, 8, and 16 Gyr after the burst, in order of increasing B − R. The error bars represent Poisson errors from sky subtractions. references within those papers). With B − R = 1.15 as a di- vision between "red" and "blue" dwarfs, we find that in the 1996 data there are 10 (12) red dwarfs rated "0" or "1" and 5 (7) blue dwarfs (where the number before the brackets is restricted to cases with colour errors < 0.2 and the number 12 Neil Trentham, R. Brent Tully and Marc A. W. Verheijen Figure 6. Colour-colour diagram for the 1999 data. The lines and symbols have the same meanings as in Figure 5, except that the filled circles now also represent galaxies rated "0". r e b m u N 12(cid:13) 10(cid:13) 8(cid:13) 6(cid:13) 4(cid:13) 2(cid:13) 0 blue galaxies red galaxies 0.5 1.0(cid:13) (cid:13)B -- R 1.5(cid:13) Figure 7. Histogram of B − R colours of the new galaxies. Both the 1996 and 1999 data are plotted. Only galaxies with an error of < 0.2 mag in B − R are included. The solid histogram represents galaxies rated "0" or "1", the shaded portion represents galaxies rated "2", and the open portion, galaxies rated "3". The division between "blue" and "red" dwarfs at B − R = 1.15, as described in the text, is shown. in brackets includes cases with large colour errors). In the 1999 data there are 5 (11) red dwarfs rated "0" and "1" and 10 (12) blue dwarfs. Most of the red dwarfs have fea- tureless morphologies, indicating that they are indeed dSph galaxies. The fraction of dwarfs that are dSph was higher in the 1996 dataset than in the 1999 dataset, suggesting that the dSphs might cluster around luminous galaxies more than dIrrs within the Ursa Major Cluster, just as Binggeli et al. (1990) found in the local Universe. By contrast with the Ursa Major Cluster, in the Virgo Cluster the vast majority of low luminosity galaxies are dwarf spheroidals (Sandage et al. 1985, Phillipps et al. 1998). CFH12K Figure 8. Projected positions of confirmed and candidate mem- bers in supergalactic coordinates. The symbols have the following meanings: filled circles -- confirmed members (from Paper I and galaxies rated "0"); open circles -- galaxies rated "1"; open squares -- galaxies rated "2"; open triangles -- galaxies rated "3". Larger symbols represent galaxies of higher luminosity. The lines are the locus of the field centers; the CFH12K field size is shown by the box in the lower right corner. The locations of the nearby 12−3 and 14−4 groups are shown. 7 DISTRIBUTION OF DWARFS IN THE CLUSTER The total area covered by the 1999 dataset was about 7.7 times the area covered by the 1996 dataset, yet we only found 1.2 times as many galaxies rated "0" or "1". It there- fore appears that galaxies are somewhat more likely to be found around luminous galaxies in the cluster than in ran- dom places in the cluster. This tendency is more marked for red dSph galaxies than for blue dIrr ones. There is a strong suggestion of a correlation between the tendency to have a substantial satellite population and parent luminosity within the 1996 dataset. The four galax- ies with the largest satellite populations are NGC 3992 (Sbc; MR = −21.93; 6 probable/possible dwarfs), NGC 3998 (E/S0; MR = −21.80; 6 probable dwarfs), NGC 4157 (Sb; MR = −21.34; 5 probable/possible dwarfs), and NGC 4100 (Sbc; MR = −21.24; 4 probable/possible dwarfs). Note that NGC 4157 was just to the north of the field centered on UGC 7176. There were four luminous Sb-bc galaxies in the 1996 dataset, only one not manifesting a significant satel- lite population (NGC 3953: Sbc; MR = −21.86; 1 probable dwarf). The three cluster fields observed in 1996 centered on a low luminosity galaxy or no known member were com- pletely devoid of new dwarf candidates. One of the two 1996 fields with the largest number of dwarfs (six) contains the NGC 3998/3990/3972 subgroup, an aggregation of three luminous galaxies (including two S0s), which may be a region of the cluster with an anoma- lously low local crossing time. One of these galaxies, NGC c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 (cid:13) 3998, is the only galaxy in Ursa Major seen to have a large globular cluster population. The 1996 fields were chosen based on prior information on the positions of galaxies so an unbiased analysis of the clustering properties within these fields is not possible (they were chosen because prior WSRT observations of the fields were available). On the other hand, the 1999 fields were chosen to simply march along the major and minor axes in the cluster, so they represent a reasonable random sampling. In Figure 8, we show the positions of the galaxies in the 1999 fields in supergalactic coordinates. It is clear from this figure that our sampling intersects two aggregations of dwarfs, one at each end of the major axis. The aggregation at the high supergalactic longitude end lies in the general vicinity of the subcondensation of early- type galaxies, NGC 4111, 4117, and 4143, in and adjacent field A30. The only other subcondensation of early type galaxies in the Ursa Major Cluster, that around NGC 3998, was found from the 1996 data to be an environment elevated in dwarfs. The aggregation at the low supergalactic longitude end might suffer serious contamination from the nearby 12-3 Group. The 12-3 Group is nearby in the sky to this end of the major axis of the cluster but at slightly higher velocity than the Ursa Major Cluster. In Paper I it is noted that it is at this border with the 12-3 Group that the Ursa Major Clus- ter is least cleanly defined. It is distinctly possible that the dwarfs in this region identified as possible/probable group members are not in the Ursa Major Cluster but, rather, are in the adjacent 12-3 Group and it will take velocity in- formation or precise distance discrimination to settle the question. We are assuming that these galaxies are members of the Ursa Major Cluster, so if they are not members the luminosity function has even a lower slope than we claim. Outside these two condensations, the clustering of dwarfs within the 1999 dataset is weak. The principal find- ing is the paucity of dwarf candidates. Along the major and minor axis strips, it is as likely to find a big, previ- ously cataloged cluster member as a dwarf galaxy proba- bly/possibly associated with the cluster. There are so few candidate dwarfs that it makes little sense to look for spa- tial correlations. We have found that the red spheroidal candidates do tend to reside near most of the luminous giant galaxies in our survey region, though they are not abundant even there. Elsewhere they are rare. The blue candidates, some with detectable HI and others with only HI upper limits, are more loosely distributed through the cluster but in remarkably small numbers. 8 DISCUSSION AND SUMMARY The following eight observations and inferences present a summary view of aspects of the Ursa Major Cluster (see also Paper I). The last six derive from the present work. For each of these new results we discuss possible reasons for why the cluster behaves the way it does in the context of the wider galaxy formation problem. (1) The Ursa Major Cluster is a diffuse cluster of 62 lu- minous galaxies with MB < −16.9 (assuming a distance d = 18.6 Mpc), all but eight of which are late-type galaxies c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The Ursa Major Cluster of Galaxies 13 (Paper I). The early-type galaxies in the cluster are lentic- ulars, with at most one elliptical, and none are extremely luminous. Five of the early-type galaxies lie in two nests at opposite ends of the cluster. (2) The mass of the Ursa Major Cluster is 5 × 1013M⊙, about one-twentieth the mass of the Virgo Cluster (the B- band mass-to-light ratio is 6 times higher for Virgo than for Ursa Major; see the introduction to Paper I) . The collapse timescale for the Ursa Major Cluster is ∼ 17 Gyrs, roughly a Hubble time (Tully 1987b). By comparison, the collapse timescale for the Virgo Cluster is ∼ 2 Gyrs, so that cluster is much more dynamically evolved. Most galaxies in the Uni- verse exist in unevolved environments like the Ursa Major Cluster. This cluster is therefore a good place to study dwarf galaxy properties in an environment akin to the field yet the density of galaxies is high enough that counting statistics have significance. (3) The luminosity function of dwarf galaxies in the Ursa Major Cluster is very flat, with a logarithmic slope α = −1.1 and an uncertainty not more than 0.2. This distribution is far shallower than the Virgo Cluster luminosity function, which has α = −2.26 ± 0.14 (Phillipps et al. 1998). Dwarf galaxies are thought to be heavily dark-matter dominated (e.g. Aaronson & Olszewski 1987, Persic & Salucci 1988, Ko- rmendy 1990, Pryor & Kormendy 1990). Cold dark matter hierarchical clustering theories of galaxy formation predict large numbers of these small dark-matter halos everywhere, a robust prediction that follows directly from the fluctuation spectrum. It appears that the expectations of the theory are met, and only met, in dynamically evolved regions like the Virgo Cluster. Hence there is apparently suppression in low density environments. Galaxies within evolved clusters have had many cross- ing times in which to interact. Gas cannot continue to be accreted onto galaxies once they enter cluster environments since the gas is outside the Roche-limit of the galaxies within the cluster potential (Shaya & Tully 1984). That gas ther- malizes and becomes the X-ray plasma. The environment in low density regions is much more benign for gas accretion onto the dwarfs. In spite of this the dwarfs are numerous in dynamically evolved environments and not in low density environments. Suppose that dwarfs that have ended up in evolved clus- ters formed very early, on top of the large scale fluctuation that grew into the cluster. Most of them have been absorbed into bigger systems but there were such large numbers of them that many survive. Over in the low density regions the entire process was slowed. The small scale perturbations that produced dwarfs collapsed over a more extended time. Supernovae and massive star formation periodically evacu- ate or rarefy the gas. If nothing else was going on, perhaps this gas could re-accumulate and continue to form stars. But something else is going on: in the low density phase the gas is being ionized by the metagalactic ultraviolet flux. The dark wells might persist but gas cannot accumulate to form stars (Klypin et al. 1999, Bullock et al. 2000). The result is a dwarf luminosity function that is steep in clusters but flat in the field. In such a scenario, the dwarf formation in Virgo hap- pens much earlier than the dwarf formation in Ursa Major; this timescale difference is consistent with what is expected given hierarchical cluster models of galaxy formation given a 14 Neil Trentham, R. Brent Tully and Marc A. W. Verheijen plausible redshift for the reionization of the Universe (Tully et al. 2001). (4) Dwarfs in the Ursa Major Cluster are as likely to be dwarf irregulars as dwarf spheroidals. This finding derives particularly from the 1999 dataset, where we observed blind fields along the minor and major axes. Many low luminos- ity galaxies therefore appear to be capable of ongoing star formation, as is the case with the giant galaxies in the clus- ter. This property is probably linked to the young dynamic age of the cluster. The galaxies in question have been rel- atively isolated all their lives and have been able to draw upon gas reservoirs over an extended time. The Ursa Ma- jor Cluster is quite different from the Virgo Cluster, where HI gas reservoirs have more frequently been depleted and the dwarf galaxies that we see there today are mostly dwarf spheroidals. (5) Most dwarfs that we detect have reasonably low sur- face brightnesses and follow the absolute magnitude vs. sur- face brightness correlation suggested by Figure 1 of Binggeli (1994; also see Tully & Verheijen 1997, Paper II). Compact dwarfs are rare. The same is true in both the local field (Binggeli, Sandage & Tammann 1988) and in richer clusters like Coma (Karachentsev et al. 1995) where the luminos- ity function of high surface brightness ellipticals is falling rapidly at faint magnitudes. (6) The dwarf galaxies seem to be clustered within the Ursa Major Cluster. In particular, there appear to be two aggre- gations, each of about twelve galaxies, at each end of the ma- jor axis. Dwarfs probably have substantial dark-matter halos which themselves are highly clustered in the context of hi- erarchical clustering models of galaxy formation. Clustering within the Ursa Major Cluster has not yet been washed out by galaxy-galaxy interactions. This circumstance is not sur- prising given the low velocity dispersion, hence long crossing time. (7) All but one of the giant galaxies targetted in the 1996 observations have a significant number of spheroidal can- didates projected within 150 kpc. Perhaps their presence manifests a small-scale version of the physical processes dis- cussed in point (3) since the local dynamical time on the mass scale of the giant galaxy is short. Alternatively, since the numbers of dwarfs are modest, these objects could be debris from ancient galaxy formation interactions (Mirabel, Dottori, & Lutz, 1992) (8) Galaxies with HI detections can be as faint as R = 19 (MR = −12). The details of the HI observations are pre- sented elsewhere (Verheijen et al. 2000). ACKNOWLEDGMENTS NT acknowledges the PPARC for financial support. Helpful discussions with M. Disney, R. Terlevich, and E. Terlevich are gratefully acknowledged. Observations were made at the Canada-France-Hawaii Telescope, which is operated by the National Research Council of Canada, the Centre National de la Recherche Scientifique de France, and the University of Hawaii. REFERENCES Aaronson M., Olszewski E., 1987, in Kormendy J., Knapp G. R., ed., IAU Symposium 117: Dark Matter in the Universe. Rei- del, Dordrecht, p. 153 Bardeen J. M., Bond J. R., Kaiser N., Szalay A. S., 1986, ApJ, 304, 15 Baugh C. M., Cole S., Frenk C. S., Lacey C. G., 1998, ApJ, 498, 504 Binggeli B., 1994, in Meylan G., Prugneil P., ed., ESO Conference and Workshop Proceedings No. 49: Dwarf Galaxies. European Space Observatory, Munich, p. 13 Binggeli B., Cameron L. M., 1991, A&A, 252, 27 Binggeli B., Sandage A., Tamman G. A., 1988, ARAA, 26, 509 Binggeli B., Tarenghi M., Sandage, A., 1990, A&A, 228, 42 Bothun G. D., Impey C. D., Malin D. F., Mould J. R., 1987, AJ, 94, 23 Bothun G. D., Schombert J. M., Impey C. D., Schneider S. E., 1990, ApJ, 360, 427 Bruzual G., Charlot S., 1993, ApJ, 405, 538 Bullock J. S., Kravstov A. V., Weinberg D. H., 2000, ApJ, sub- mitted, (astro-ph/00002214) Coleman G. D., Wu C-C., Weedman D. W., 1980, ApJS, 43, 393 Cowie L. L. Songaila A., Hu E. M., Cohen J. G., 1996, AJ, 112, 839 Cuillandre J.-C., Luppino G., Starr B., Isani S., 1999, CFHT Bulletin 40 Davies J. I., Phillipps S., Disney M. J., 1988, MNRAS, 231, 69p Diaferio A., Kauffmann G., Colberg J. M., White S. D. M., 1999, MNRAS, 307, 537 Disney M. J., 1976, Nat, 263, 573 Ellis R. S., Colless M., Broadhurst T., Heyl J., Glazebrook K., 1996, MNRAS, 280, 235 Ferguson H. C., Sandage A., 1991, AJ, 101, 765 Impey C., Bothun G., Malin D., 1988, ApJ, 330, 634 Jarvis J. F., Tyson J. A., 1981, AJ, 86, 476 Karachentsev I. D., Karachentseva V. E., Richter G. M., Vennik J. A., 1995, A&A, 296, 643 Klypin A., Kravtsov A., Valenzuela O., Prada F., 1999, ApJ, 522, 82 Kormendy J., 1990, in Kron R. G., ed., The Edwin Hubble Cen- tennial Symposium: The Evolution of the Universe of Galax- ies. Astronomical Society of the Pacific, San Francisco, p. 33 Kormendy J., Djorgovski S., 1989, ARAA, 27, 235 Landolt A. U., 1992, AJ, 104, 340 Lin H., Kirshner R. P., Shectman S. A., Landy S. D., Oemler A., Tucker D. L., Schechter P. L., 1996, ApJ, 464, 60 Lynds R., Tolstoy E., O'Neil E., Hunter D. A., 1998, AJ, 116, 146 Marzke R. O, Geller M. J., Huchra J. P., Corwin H. G., 1994, AJ, 108, 437 Metzger M. R., Luppino G. A., Miyazaki S., 1995, BAAS, 187, 73.05 Mirabel I. F., Dottori H., Lutz D., 1992, A&A, 256, L19 Persic M., Salucci P., 1988, MNRAS, 234, 131 Phillipps S., Parker Q. A., Schwartzenberg J. M., Jones J. B., 1998, ApJ, 493, L59 Press W. H., Schechter P., 1974, ApJ, 187, 425 Pryor C., Kormendy J., 1990, AJ, 100, 127 Pustilnik S., Engles D., Ugryumov V., Lipovetsky V. A., Hagen H. J., Kniazev A. Y., Izotov Y. I., Richter G., 1999, A&AS, 137, 299 Salpeter E. E., 1955, ApJ, 121, 161 Sandage A., Binggeli B., Tammann G. A., 1985, AJ, 90, 1759 Schechter P., 1976, ApJ, 203, 297 Schlegel D. J., Finkbeiner D. P., Davis M., 1998, ApJ, 500, 525 Schneider D. P., Schmidt M., Gunn J. E., AJ, 107, 1245 Shaya E. J., Tully R. B., 1984, ApJ, 281, 56 Somerville R. S., Primack J. R., 1999, MNRAS, 310, 1087 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The Ursa Major Cluster of Galaxies 15 Telles E., Melnick J., Terlevich R. J., 1997, MNRAS, 288, 78 Trentham N., 1998, MNRAS, 295, 360 Trentham N., Tully R. B., Verheijen M. A. W., 2001, MNRAS, submitted Tully R. B., 1987a, Nearby Galaxies Catalog, Cambridge Univer- sity Press, Tully R. B., 1987b, ApJ, 321, 280 Tully R. B., 1988, AJ, 96, 73 Tully R. B., Pierce M J., 2000, ApJ, 533, 744 Tully R. B., Somerville, R. S., Trentham N., & Verheijen M. A. W., 2001, ApJ, to be submitted Tully R. B., Verheijen M. A. W., 1997, ApJ, 484, 145 (Paper II) Tully R. B., Verheijen M. A. W., Pierce M. J., Huang J. -S., Wain- scoat R. J., 1996, AJ, 112, 2471 (Paper I) Valdes F., 1982, Proc. SPIE, 331, 465 Valdes F., 1989, in Grosbol P. J., Murtagh F., Warmels R. H., ed., ESO Conference and Workshop Proceedings No. 31: Pro- ceedings of the 1st ESO/St-ECF Data Analysis Workshop. European Space Observatory, Munich, p. 35 van den Bergh S., 1992, A&A, 264, 75 van den Bergh S., 2000, PASP, 112, 529 Verheijen M. A. W., 1998, PhD Thesis, University of Groningen Verheijen M., Trentham N., Tully B., Zwaan M., 2000, in Kraan- Korteweg R. C., Henning P. A., Andernach H., eds, Mapping the Hidden Universe: The Universe Behind the Milky Way - The Universe in HI. ASP, San Francisco, in press (astro- ph/0006122) Verheijen M., Trentham N., Tully B., Zwaan M., 2001, in prepa- ration White S. D. M., Rees M. J., 1978, MNRAS, 183, 321 Whiting A. B., Hau G. K. T., Irwin M., 1999, AJ, 118, 2767 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 This figure "fig3p1.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0103039v1 This figure "fig3p2.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0103039v1 This figure "fig3p3.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0103039v1 This figure "fig3p4.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0103039v1 This figure "fig3p5.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0103039v1 This figure "fig3p6.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0103039v1 This figure "fig3p7.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0103039v1
astro-ph/0610434
1
0610
2006-10-13T22:40:00
Revealing the High Energy Emission from the Obscured Seyfert Galaxy MCG -5-23-16 with Suzaku
[ "astro-ph" ]
We report on a 100 ks Suzaku observation of the bright, nearby (z=0.008486) Seyfert 1.9 galaxy MCG -5-23-16. The broad-band (0.4-100 keV) X-ray spectrum allows us to determine the nature of the high energy emission with little ambiguity. The X-ray continuum consists of a cutoff power-law of photon index $\Gamma=1.9$, absorbed through Compton-thin matter of column density $N_{\rm H}=1.6\times10^{22}$ cm$^{-2}$. A soft excess is observed below 1 keV and is likely a combination of emission from scattered continuum photons and distant photoionized gas. The iron K line profile is complex, showing narrow neutral iron K$\alpha$ and K$\beta$ emission, as well as a broad line which can be modeled by a moderately inclined accretion disk. The line profile shows either the disk is truncated at a few tens of gravitational radii, or the disk emissivity profile is relatively flat. A strong Compton reflection component is detected above 10 keV, which is best modeled by a combination of reflection off distant matter and the accretion disk. The reflection component does not appear to vary. The overall picture is that this Seyfert 1.9 galaxy is viewed at moderate (50 degrees) inclination through Compton-thin matter at the edge of a Compton-thick torus covering $2\pi$ steradians, consistent with unified models.
astro-ph
astro-ph
PASJ: Publ. Astron. Soc. Japan , 1 -- ??, c(cid:13) 2021. Astronomical Society of Japan. Revealing the High Energy Emission from the Obscured Seyfert Galaxy MCG -5-23-16 with Suzaku 6 0 0 2 t c O 3 1 1 v 4 3 4 0 1 6 0 / h p - o r t s a : v i X r a James N. Reeves1,2 Hisamitsu Awaki3 Gulab C. Dewangan4 Andy C. Fabian5 Yasushi Fukazawa6 Luigi Gallo7,8 Richard Griffiths4 Hajime Inoue8 Hideyo Kunieda8,9 Alex Markowitz1 Giovanni Miniutti5 Tsunefumi Mizuno6 Richard Mushotzky1 Takashi Okajima1,2 Andy Ptak1,2 Tadayuki Takahashi8 Yuichi Terashima3,8 Masayoshi Ushio8 Shin Watanabe8 Tomonori Yamasaki6 1Exploration of the Universe Division, Code 662, NASA Goddard Space Flight Center, Greenbelt Road, Makoto Yamauchi10 and Tahir Yaqoob1,2 Greenbelt, MD 20771, USA [email protected] 2Department of Physics and Astronomy, Johns Hopkins University, 3400 N Charles Street, 3Department of Physics, Ehime University, Matsuyama 790-8577, Japan 4Department of Physics, Carnegie Mellon University, 5000 Forbes Avenue, Pittsburgh, Baltimore, MD 21218, USA 5Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge, PA 15213, USA 6Department of Physics, Hiroshima University, 1-3-1 Kagamiyama, Higashi-Hiroshima, CB3 0HA, UK Hiroshima 739-8526, Japan 7Max-Planck-Institut fur extraterrestrische Physik, Postfach 1312, Garching, Germany 8Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, Yoshinodai 3-1-1, 9Department of Physics, Nagoya University, Furo -- cho, Chikusa, Nagoya 464-8602, Japan 10Department of Applied Physics, University of Miyazaki, 1-1, Gakuen-Kibanadai-Nishi, Sagamihara, Kanagawa 229-8510, Japan Miyazaki 889-2192, Japan (Received 2006 July 27; accepted 2006 September 18) Abstract We report on a 100 ks Suzaku observation of the bright, nearby (z = 0.008486) Seyfert 1.9 galaxy MCG -5-23-16. The broad-band (0.4 -- 100 keV) X-ray spectrum allows us to determine the nature of the high energy emission with little ambiguity. The X-ray continuum consists of a cutoff power-law of photon index Γ = 1.9, absorbed through Compton-thin matter of column density NH = 1.6 × 1022 cm−2. A soft excess is observed below 1 keV and is likely a combination of emission from scattered continuum photons and distant photoionized gas. The iron K line profile is complex, showing narrow neutral iron Kα and Kβ emission, as well as a broad line which can be modeled by a moderately inclined accretion disk. The line profile shows either the disk is truncated at a few tens of gravitational radii, or the disk emissivity profile is relatively flat. A strong Compton reflection component is detected above 10 keV, which is best modeled by a combination of reflection off distant matter and the accretion disk. The reflection component does not appear to vary. The overall picture is that this Seyfert 1.9 galaxy is viewed at moderate (∼ 50◦) inclination through Compton-thin matter at the edge of a Compton-thick torus covering ∼ 2π steradians, consistent with unified models. Key words: galaxies: individual (MCG -5-23-16); galaxies: active; galaxies: Seyfert; X-rays: galaxies 1. Introduction Determining the origin of the iron K emission line is one of the fundamental issues in high energy research on Active Galactic Nuclei, as it is regarded as the most di- rect probe available of the inner accretion disk and black hole. Indeed the iron line diagnostic in AGN first became important during the Ginga era, showing that the 6.4 keV iron Kα emission line was common amongst Seyfert galax- ies (Pounds et al. 1990). The associated reflection hump above 10 keV, produced by Compton down-scattering of higher energy photons, showed the iron line emission arises from Compton-thick material, possibly the accretion disk (Lightman & White 1988; George & Fabian 1991). The higher (CCD) resolution spectra available with ASCA ap- peared to indicate that the iron line profiles were broad and asymmetrically skewed (to lower energies), which was interpreted as evidence that the majority of the line emission originated from the inner accretion disk around the massive black hole (Tanaka et al. 1995; Nandra et al. 1997; Reynolds 1997). The picture emerging from the study of the iron K line 2 Reeves et al. [Vol. , with XMM-Newton and Chandra is much more complex. The presence of a narrower 6.4 keV iron emission compo- nent, from more distant matter (e.g. the outer disk, BLR or the molecular torus) is commonplace in many type I AGN (Yaqoob & Padmanabhan 2004; Page et al. 2004). In contrast the broad, relativistic component of the iron line profile appears to be weaker than anticipated and in some cases may be absent altogether, e.g. NGC 5548 (Pounds et al. 2003a); NGC 4151 (Schurch et al. 2003). Furthermore, it is possible that complex absorption in some objects may also effect the modeling of the iron K- line and reduce its strength; e.g. in NGC 3783 (Reeves et al. 2003) and NGC 3516 (Turner et al. 2005). The pos- sible presence of the reflection component hardening the spectrum towards higher energies (George & Fabian 1991) can also complicate fitting the broad iron K line, if band- pass above 10 keV is not available. This is where new ob- servations with Suzaku can provide an important break- through, by determining the underlying AGN continuum emission over a wide bandpass (e.g. from 0.3 keV to > 100 keV), thereby resolving the ambiguities present in fitting the iron K-shell band. MCG -5-23-16 is an X-ray bright, nearby (z = 0.008486 or 36 Mpc, Wegner et al. 2003) AGN that is classed op- tically as a Seyfert 1.9 galaxy (Veron et al. 1980) and is known to possess moderate absorption in the soft X- ray band (Mushotzky 1982). It is one of the brightest Seyfert galaxies in the 2-10 keV band, with a historically typical flux of 7 − 10 × 10−11 erg cm−2 s−1 (Mattson & Weaver 2004), which makes it an ideal object to study in the iron K band and at high energies to measure any Compton reflection component. Indeed of the objects that now show evidence for a broad redshifted iron K line, MCG -5-23-16 appears to be one of the more ro- bust examples from XMM-Newton data (Dewangan et al. 2003; Balestra et al. 2004) or with ASCA (Weaver et al. 1997). Here we report on a 100 ks observation of MCG -5-23- 16, performed in December 2005 with Suzaku (Mitsuda et al. 2006). In Section 2, the Suzaku observations of MCG - 5-23-16 are outlined, whilst in Section 3 the detailed mod- eling of the time-averaged spectrum is performed. In Section 4 the variability of the iron line and reflection component are compared while in Section 5, modeling of the reflection component is discussed. 2. The Suzaku Observation of MCG-5-23-16 MCG -5-23-16 was observed by Suzaku between 7-10 December 2001 with a total duration of 220 ks (see Table 1 for a summary of observations). Events files from ver- sion 0.7 of the Suzaku pipeline processing were used. 1 1 Version 0 processing is a simplified processing applied to the Suzaku data obtained during the SWG phase, for the purpose of establishing the detector calibration as quick as possible. Some processes that are not critical for most of the initial scientific studies, e.g., aspect correction, fine tuning of the event time tagging of the XIS data, are skipped in version 0 processing, and hence, the quality of the products is limited in these directions, Fig. 1. The 2-10 keV XIS 3 lightcurve of MCG -5-23-16, shown as black circles. The observation started on 7 December 2005, with a total exposure time (after screening) of ∼ 100 ks. The Suzaku lightcurve has been binned into orbital length bins of 5760s. Overlaid in red is the XMM-Newton EPIC-pn lightcurve, which is simultaneous for part of the ob- servation. All events files were further screened within xselect to exclude data within the SAA (South Atlantic Anomaly) as well as excluding data with an Earth elevation an- gle (ELV) < 5 degrees. Data with Earth day-time ele- vation angles (DYE ELV) less than 20 degrees were also excluded. Furthermore data within 256s of the SAA were excluded from the XIS and within 500s of the SAA for the HXD (using the T SAA HXD parameter within the house-keeping files). Cut-off rigidity (COR) criteria of > 8 GeV/c for the HXD data and > 6 GeV/c for the XIS were used. 2.1. XIS Data Analysis For the XIS, only good events with grades 0,2,3,4 and 6 were used, while hot and flickering pixels were removed from the XIS images using the cleansis script. Time intervals effected by telemetry saturation were also re- moved. The XIS was set to normal clocking mode and the data format was either in the 3 × 3 or 5 × 5 edit modes (corresponding to 9 and 25 pixel pulse height arrays re- spectively that are telemetered to the ground). The XIS pulse height data for each X-ray event were converted to PI (Pulse Invariant) channels using the xispi software version 2005/12/26 and CTI parameters were used from 2006/05/22. The 4 XIS source spectra (see Koyama et al. (2006) for details of the XIS) were extracted from circular source re- gions of 4.34′ (6mm) radius centered on the source, which was observed off-axis in the HXD nominal pointing posi- tion. Background spectra were extracted from three 2.5′ circles offset from the source region. The source and back- ground extraction regions were also chosen to avoid the calibration sources on the corners of the CCD chips. XIS compared with the official data supplied to the guest observers. As of 2006 July, version 0.7 is the latest, where the absolute energy scale of ∼ 5 eV is achieved for the XIS data. No. ] Revealing the High Energy Emission From MCG -5-23-16 3 Table 1. Summary of Observations TSTART (UT) Instrument 2005/12/07 22:59:49 Suzaku/XIS 2005/12/07 22:40:00 Suzaku/PIN Suzaku/GSO 2005/12/07 22:40:00 XMM-Newton/PN 2005/12/08 21:11:33 2005/12/08 17:41:30 Chandra/HETG 2005/12/09 20:52:11 Chandra/HETG 2005/12/09 13:16:00 RXTE/PCA Swift/XRT 2005/12/09 00:03:00 TSTOP (UT) Exposure (ks) 2005/12/10 11:48:54 2005/12/10 11:50:50 2005/12/10 11:50:50 2005/12/10 06:57:51 2005/12/09 02:33:59 2005/12/10 03:00:10 2005/12/09 20:30:00 2005/12/09 09:54:57 98.1 71.4 18.3 96.2 49.5 12.8 10.1 response files (rmfs) provided by the instrument teams were used (dated 2006/02/13), while ancillary response files (arfs) appropriate for the HXD nominal pointing were chosen dated as of 2006/04/15, matching the size (6mm) of the source extraction regions used. For the front illu- minated XIS chips (XIS 0,2,3) data over the energy range 0.6 -- 10 keV were used, while for the softer back-illuminated XIS 1 chip data from 0.4 -- 8 keV were used. Data around the CCD Si K edge from 1.7 -- 1.95 keV were ignored in all 4 XIS chips, due to uncertain calibration at this energy at the time of writing. A total source exposure (after screening) of 98.1 ks was obtained for the 4 XIS chips. The 2-10 keV source lightcurve for the XIS 3 is shown in Figure 1. The XIS spectra were corrected for the hydrocarbon (C24H38O4) contamination on the optical blocking filter, by including an extra absorption column due to Carbon and Oxygen in all the spectral fits. The column densi- ties for each detector were calculated based on the date of the observation and the off-axis position of the source. The Carbon column densities (NC) used were 1.64 × 1018, 2.18 ×1018, 2.82 ×1018 and 4.40 ×1018 atoms cm−2 for XIS 0,1,2,3 respectively, with the ratio of C/O column densi- ties set to 6. The additional soft X-ray absorption due to the hydrocarbon contamination was included as a fixed spectral component using the varabs absorption model in all the spectral fits. The net source count rates obtained for the 4 XIS detec- tors are 2.843 ± 0.006 counts s−1, 3.081 ± 0.006 counts s−1, 3.403 ± 0.007 counts s−1 and 3.316 ± 0.007 counts s−1 for XIS 0 -- 3 respectively, with background typically 1% of the source rate. Given the large number of source counts available in the observation, the XIS source spectra were binned to a minimum of 200 counts per bin to enable the use of the χ2 minimization technique. 2.2. HXD Data Analysis 2.2.1. HXD/PIN The source spectrum was extracted from the cleaned HXD/PIN events files, processed with the screening cri- teria described above. The HXD/PIN instrumental back- ground spectrum was generated from a time dependent model provided by the HXD instrument team. The model utilized the count rate of upper discriminators as the mea- sure of cosmic-ray flux that passes through the silicon PIN diode and provide background spectra based upon a database of non-X-ray background observations made by the PIN diode to date (Kokubun et al. 2006). Both the source and background spectra were made with iden- tical GTIs (good time intervals) and the source expo- sure was corrected for detector deadtime (which is < 5%). A detailed description of the PIN detector deadtime is given in Kokubun et al. (2006). The net exposure time of the PIN source spectrum was 71.4 ks after deadtime correction. Note that the background spectral model was generated with 10× the actual background count rate in order to minimize the photon noise on the background; this has been accounted for by increasing the effective exposure time of the background spectra by a factor of ×10. The HXD/PIN response file dated 2006/08/14 for the HXD nominal position was used for these spectral fits. Instrumental and performance details of the HXD are discussed in Takahashi et al. (2006) and Kokubun et al. (2006). In addition a spectrum of the cosmic X-ray back- ground (CXB) (Boldt 1987; Gruber et al. 1999) was also simulated with the HXD/PIN. The form of the CXB was taken as 9.0 × 10−9 × (E/3 keV)−0.29 × e(−E/40 keV) erg cm−2 s−1 sr−1 keV−1. When normalized to the field of view of the HXD/PIN instrument the effective flux of the CXB component is 9.0 × 10−12 erg cm−2 s−1 in the 12 -- 50 keV band. Note the flux of MCG -5-23- 16 measured by the HXD over the same band is 1.6 × 10−10 erg cm−2 s−1, i.e. the CXB component represents only 6% of the net source flux measured by the HXD/PIN. Note that there is some uncertainty in the absolute flux level of the CXB component measured between missions, for instance Churazov et al. (2006) find the CXB normal- ization from Integral to be about 10% higher than mea- sured by Gruber et al. (1999) from the HEAO-1 data. However this level of uncertainty is much smaller than the net source flux of MCG -5-23-16, at the level of < 1%. Furthermore the spatial fluctuation of the CXB also has a negligible effect (< 1% of the net source flux), for instance Kushino et al. (2002) measure a 6.5 ± 0.6% fluctuation based on the 2 -- 10 keV ASCA GIS data, which has a total field of view similar to Suzaku HXD/PIN. The HXD/PIN spectrum is shown in Figure 2, plot- ted from 12 -- 50 keV. At present we exclude the HXD/PIN data above 50 keV, as a detailed study of the background systematics for HXD/PIN is on-going above this energy. Figure 2 shows the total (source + observed background) 4 Reeves et al. [Vol. , Fig. 3. The four Suzaku XIS spectra (XIS 0, black; XIS 1, red; XIS 2; green; XIS 3; blue) and HXD/PIN (light blue stars) over the 2-50 keV range plotted as a ratio to an ab- sorbed power-law of photon index Γ = 1.8. A 15% cross nor- malization factor has been accounted for between HXD/PIN and the XIS (the HXD/PIN is thought to be 13 -- 15% higher with the current calibration). The cross-normalization be- tween the 4 XIS detectors above 2 keV agrees within 5% and the photon index is consistent within Γ = ±0.05 for the 4 XIS spectra. There is an excess of counts above 12 keV in the HXD/PIN compared to the XIS, most likely due to a Compton reflection hump. 5-23-16 observation. Abell 1795 was not detected in the HXD/GSO and is therefore suitable for use as a back- ground template given the close proximity in time of the two observations. Both the GSO background from Abell 1795 and the MCG -5-23-16 source spectra were extracted from orbits that contained no SAA passage, which have the most reliable (and least variable) GSO background. Matching orbits (i.e. at the same time of day) were also chosen between the MCG -5-23-16 and Abell 1795 observations. After applying this additional screening, the net exposure times for the GSO source and background spectra were then 18.3 ks and 28.6 ks respec- tively. The source was detected at the 15σ statistical level over the GSO background in the 50-100 keV band and at a level of 5.5% over the detector background. As a study of the background systematics of the GSO is still on-going, unless otherwise stated we do not include the GSO data in the subsequent spectral fits. Nonetheless a plot of the GSO detection between 50 -- 100 keV and of the HXD/PIN and XIS spectra is shown in Figure 4, compared to an ab- sorbed power-law spectrum (see description later). Above 100 keV the GSO background systematics are more uncer- tain and subject to further study, so a reliable detection of this source above this energy cannot be made at the present time. 2.3. Simultaneous Observatories Observations With Other Simultaneous observations were also conducted with XMM-Newton, Chandra (HETG), RXTE and Swift cov- ering part of the Suzaku observation. The details of the observations are shown in Table 1. Figure 1 shows the Fig. 2. HXD/PIN spectra for the Suzaku observation of MCG -5-23-16. The upper panel shows the total spectrum (source plus background, black). The points in red and blue show two independent background models for HXD/PIN, the red points are the ones adopted in the paper. The two back- grounds show very good agreement. The net source counts (i.e. total minus model background) are shown in green. Note the source is detected in each data bin at a minimum of 10σ above the background. The lower panel shows the ratio of the net source spectrum to the total spectrum, which is > 20% over the 12 -- 50 keV range. spectrum and two independent background model spectra provided by the HXD instrument team (the red and blue points), both of which include the instrumental (non X- ray) background and the contribution from the CXB. The net source spectrum is plotted (green points), which shows the total spectrum minus the background model spec- trum (the red points are the background model adopted in this paper). The agreement between the two independent background models is excellent, the spectral fits derived for MCG -5-23-16 produces consistent results (within the statistical errors quoted in the paper) for each background model. The lower panel of Figure 2 shows the ratio of net source counts to the total observed PIN counts. Thus the net source counts are > 20% of the total counts for each bin from 12 -- 50 keV, while the source is detected at > 10σ above the background in each bin in the spectrum. The resulting net PIN source count rate from 12 -- 50 keV was 0.454 ±0.003 counts s−1, compared to the PIN background rate of 0.494 ± 0.001 counts s−1. 2.2.2. HXD/GSO For the HXD/GSO, background spectra were extracted from a Suzaku observation of the cluster Abell 1795 which occurred from December 10 -- 13, 2005, just after the MCG - No. ] Revealing the High Energy Emission From MCG -5-23-16 5 lightcurve from Suzaku XIS 3 compared to the XMM- Newton EPIC-pn observation, with the 100 ks XMM- Newton observation overlapping with part of the Suzaku observation. The pattern=0 data only were used for the EPIC-pn. A more detailed description of the analysis of the simultaneous observations will appear in a subsequent paper (Braito et al. 2006, in preparation); the purpose of this paper is to describe the Suzaku observation especially the properties of the iron K line and the hard X-ray emis- sion and reflection component above 10 keV. Nonetheless we refer below to the Chandra HETG data for the mea- surement of the width of the iron Kα line core, to XMM- Newton to check the consistency of the iron K line profile with Suzaku and to RXTE as an independent check of the reflection component. 3. Broad-Band X-ray Spectral Analysis 3.1. Initial Spectral Fitting We first concentrate our analysis on the time-averaged MCG -5-23-16 spectrum, from the whole Suzaku obser- vation. The XIS and HXD/PIN background subtracted spectra were fitted using xspec v11.3, including data over the energy range 0.4 -- 50 keV. A Galactic absorption col- umn of NH = 8.0 × 1020 cm−2 (Dickey & Lockman 1990) was included in all the fits and spectral parameters are quoted in the rest-frame of MCG -5-23-16 at z = 0.008486. All errors are quoted at 90% confidence for one interest- ing parameter (corresponding to ∆χ2 = 2.7) unless other- wise stated. A cosmology of H0 = 70 km s−1 Mpc−1 and Λ0 = 0.73 is assumed throughout. Initially we fitted the 4 XIS spectra and HXD data using a simple absorbed power-law model in the 2 -- 5 keV band in order for the continuum parameters not to be affected by the iron Kα line at 6.4 keV, the reflection component at high energies, as well as any soft X-ray spectral complex- ity below 1 keV. The continuum was fitted with a photon index Γ ∼ 1.8 and an absorption column of 1.6×1022 cm−2. The relative normalization of the HXD/PIN was initially fixed at 15% above that of the XIS, as has been found to date from calibration observations of the Crab (Kokubun et al. 2006). Figure 3 shows the spectrum and data/model ratio of the 4 XIS spectra as well as the HXD/PIN data between 2 -- 50 keV against the absorbed power-law contin- uum. The photon indices of the 4 XIS spectra are all in good agreement to within ±0.05 of each other; these are Γ = 1.84 ± 0.03 (for XIS 0), Γ = 1.79 ± 0.04 (XIS 1), Γ = 1.82 ± 0.03 (XIS 2) and Γ = 1.82 ± 0.03 (XIS 3). The cross normalization between all 4 XIS detectors is also in good agreement to ±5%. An excess of flux in the HXD/PIN compared to the XIS spectra is clearly appar- ent. Due to the good agreement between the XIS spectra, data from the three front illuminated XIS 0, XIS 2 and XIS 3 chips were co-added (hereafter referred to as XIS FI) in order to maximize signal to noise. We do not combine the XIS 1 which is a back illuminated chip and its re- sponse is better suited to the soft X-ray spectrum, with lower effective area at high energies. Figure 4 shows the Fig. 4. The broad-band (0.4-100 keV) Suzaku spectrum of MCG -5-23-16. The upper panel shows the data, plotted against an absorbed power-law model of photon index Γ = 1.8 (solid line) and column density 1.6 × 1022 cm−2, fitted over the 2-5 keV band. The lower panel shows the data/model ra- tio residuals to this power-law fit. The data from XIS FI chips are shown in black, the XIS 1 in blue, the HXD/PIN as red stars and the HXD/GSO as green squares. Clear deviations in the iron K-shell band are apparent between 6-7 keV, while an excess of counts (due to a soft excess) is present below 1 keV. A hard excess is seen in the HXD spectrum above 10 keV, due to the presence of a reflection component in this object. spectrum and data/model residuals to the entire Suzaku spectrum from XIS FI, XIS 1, HXD/PIN and HXD/GSO over the 0.4-100 keV range against the simple absorbed power-law model. Note we include the GSO datapoint here for comparison, but it is not used in the spectral fit- ting. An excess of counts above the power-law continuum is seen above 10 keV likely due to Compton reflection in the X-ray spectrum, while the spectrum appears to turn over at the very highest energies, perhaps due to a high energy exponential cut-off. At lower energies a soft excess is seen below 1 keV (especially in the softer XIS 1 instru- ment), which may be due to an unabsorbed component in scattered emission associated with this Seyfert 1.9 galaxy. 3.2. The Baseline Continuum Model We then constructed a baseline model which can fit the spectrum of MCG -5-23-16. Initially we concen- trated on the joint fit between XIS FI and HXD/PIN from 0.4 -- 50 keV. For the continuum, we modeled the spectrum with an absorbed power-law component includ- ing an exponential cut-off at high energies of the form, NPLE−Γ × e−E/Ecut ; where NPL is the normalization of the power-law (in photons cm−2 s−1 keV−1 at 1 keV), E is the energy (in keV), Γ is the photon index and Ecut is the e-folding cut-off energy (in keV). This primary power-law continuum is absorbed by two layers of neutral gas; an absorbing layer at the red- shift of MCG -5-23-16 and the Galactic absorption col- umn of 8 × 1020 cm−2 (Dickey & Lockman 1990). A sec- ond soft X-ray power-law component was included, ab- 6 Reeves et al. [Vol. , sorbed by only the Galactic column, in order to model the soft excess. Abundances were set those of Wilms et al. (2000) using the cross-sections of Balucinska-Church & McCammon (1992). Furthermore a reflection compo- nent produced by Compton down-scattering of X-rays off neutral material was included, using the pexrav model within xspec (Magdziarz & Zdziarski 1995). The re- flection component is absorbed by the Galactic column only. The continuum parameters (Γ, NPL and Ecut) were tied to those of the incident cut-off power-law. The in- clination (θ) of the reflector was set to 45◦, while the Fe abundance (AFe) was allowed to vary. The solid an- gle (Ω) subtended by the Compton reflector was deter- mined by the parameter R = Ω/2π. The best-fit param- eters are shown in Table 2. The primary power-law has a photon index of Γ = 1.95 ± 0.03 after the inclusion of the reflection component, with a high energy cut-off of Ecut > 170 keV, absorbed by a neutral column density of NH = 1.65 ± 0.03 × 1022 cm−2. The soft, unabsorbed power-law required to fit the soft excess has a steeper pho- ton index of Γ = 3.1 ± 0.3. The properties of the iron K line and the reflection component are described below. 3.3. The Iron K Line Profile The data/model residuals of the XIS FI spectrum in the iron K band to this continuum model are shown in Figure 5 (without the inclusion of the reflection compo- nent). For comparison, the XMM-Newton EPIC-pn data are also plotted, which shows that the iron K line profiles obtained by XMM-Newton and Suzaku are in excellent agreement. The data from both missions show a strong iron Kα core at 6.4 keV, an excess of counts both red- wards and blue-wards of 6.4 keV, a weak peak at 7 keV due to the iron Kβ line and a drop at 7.1 keV, likely due to the neutral Fe K edge associated with the Compton reflector. The iron K line profile was then fitted in several steps. Firstly Compton reflection was included, as described above. The resultant line profile to the Suzaku data can be seen in Figure 6, panel (a). Then the narrow core to the Kα line was added to the model, while a narrow Kβ line was also included, the line energy of the Kβ line was fixed at 7.056 keV, with the line flux set equal to 12% of the Kα as expected for neutral iron. For complete- ness, we also add a small Compton shoulder to the iron Kα line, represented by a narrow Gaussian centered at 6.3 keV, with normalization fixed to 20% of the Kα flux, expected if the emission originates from Compton-thick material (Matt 2002). After the inclusion of the narrow lines, a clear excess of counts in the Suzaku data can be seen around 6.4 keV (Figure 6, panel b). This excess can modeled by either a broad Gaussian line or a diskline pro- file from an accretion disk (Fabian et al. 1989). Figure 6(c) shows the residuals in the iron K band after fitting the iron K line with the narrow Kα and Kβ lines, as well as a broad iron line component. The fit statistic obtained is good considering the high statistical quality of the data (χ2/dof = 1983/1879 for the disk-line model, where dof is the number of degrees of freedom). Fig. 5. The iron line profile of MCG -5-23-16, plotted as a ratio against a power-law of photon index Γ = 1.8. The data from Suzaku XIS FI is shown in black and for comparison the blue circles show the data from the XMM-Newton EPIC-pn observation. Both observations show a narrow iron Kα core at 6.4 keV, a red-wing below 6.4 keV, a peak at 7.05 keV due to Fe Kβ. The edge at 7.1 keV is due to the reflection component. 3.3.1. The Narrow Kα Line The narrow Kα core has an equivalent width of 70 ± 6 eV, centered at 6.391 ±0.007 keV, with a measured width of σ = 51 ± 12 eV. The 55Fe calibration source located on the corners of the XIS chips can be used as an accurate calibrator of the energy and intrinsic width of the iron line. The 55Fe source produces lines from Mn Kα (Kα1 & Kα2 at 5.899 keV and 5.888 keV respectively with a branching ratio of 2:1). From measuring the lines in the calibration source, we find that the line energy is shifted redwards by 9 ± 2 eV, while there is some residual width (after the instrumental response function has been accounted for) in the calibration lines of σ = 41 ± 2 eV. Therefore the intrinsic width (σint) of the iron Kα line is simply σ2 int = σ2 cal (where σmeas is the measured width and σcal meas is the residual width of the calibration lines). − σ2 The intrinsic width is thus formally consistent with being unresolved, while a statistical upper limit can be placed on the width of σint < 48 eV, corresponding to a velocity width of σv < 2200 km s−1. The corrected en- ergy of the narrow iron line is then 6.400 ± 0.007 keV, consistent with a superposition of the neutral Kα1 and Kα2 iron lines at 6.404 keV and 6.391 keV respectively. In comparison, the line width and energy obtained from the simultaneous Chandra/HETG observation are σ < 43 eV (σv < 2000 km s−1) and 6.405 ± 0.015 keV (Braito et al. 2006) which is in excellent agreement. Note that the residual broadening in the XIS response function has a negligible effect on the broad line parameters determined below. 3.3.2. The Broad Iron Line When modeled by a Gaussian, the broad component of the line is also centered near 6.4 keV (E = 6.43 ±0.10 keV), with an equivalent width of 62 ± 17 eV and an intrinsic No. ] Revealing the High Energy Emission From MCG -5-23-16 7 Table 2. Spectral Fit Parameters. Note a normalization of power-law in units 10−2 photons cm−2 s−1 keV−1 at 1 keV; b X-ray flux in units of 10−11 erg cm−2 s−1; c column density in units of 1022 cm−2; d normalization (flux) of iron line in units 10−5 photons cm−2 s−1; e inner disk radius in gravitational radii. f donates parameter is fixed in fit. b c a a Parameter Γ NPL Ecut (keV) b Flux2−10 Flux15−100 NH Γ NPL R Abund (Fe) E (keV) σ (eV) d Nline EW (eV) E (keV) σ (keV) Nline EW (eV) E (keV) d Continuum Scattered Reflection Narrow Gaussian Broad Gaussian or Diskline (XMM+Suzaku) Rin e θ Nline EW (eV) d Mean 1.95 ± 0.03 3.2 ± 0.1 High Low 1.92 ± 0.03 3.5 ± 0.1 1.93 ± 0.03 2.6 ± 0.1 > 170 8.76 19.0 1.65 ± 0.03 3.1 ± 0.3 1.5 ± 0.2 × 10−2 1.1 ± 0.2 0.40 ± 0.13 200f 9.91 20.0 200f 7.12 14.1 1.2 ± 0.2 1.3 ± 0.2 0.5f 0.5f 6.400 ± 0.007 6.40 ± 0.01 6.40 ± 0.01 10f 6.5 ± 2.0 65 ± 19 6.4 ± 0.1 0.4f 5.6+3.1 −3.6 53 10f 5.4+1.7 −2.6 91+28 −44 6.4 ± 0.1 0.4f 6.8+3.4 −2.9 92 < 48 6.5 ± 0.5 70 ± 6 6.39 ± 0.10 0.44 ± 0.12 6.1 ± 1.6 62 ± 17 6.40f 37+25 −10 53+7 −9 5.1 ± 0.9 60 ± 11 ◦, the inner radius of the disk is 26+35 width of σ = 440 ± 120 eV or σv ∼ 20000 km s−1. This large velocity indicates that the broad line must originate from close to the black hole in MCG -5-23-16. Therefore the broad iron line was modeled with an emission profile from an accretion disk around a Schwarzschild black hole (diskline, Fabian et al. (1989)). The outer radius of the diskline was fixed to 400Rg (where Rg = GM/c2), while it was assumed that the iron K emission originates from neutral matter at 6.4 keV. The disk emissivity is parame- terized in the form r−q, where q is the emissivity index and r is the disk radius. Initially the emissivity was fixed at q = 3. The disk inclination angle is then constrained to be 50+32 −8 Rg, while the −10 line EW is 63 ±16 eV. If the constraint on the disk emissiv- ity is relaxed, then a disk inner radius of 6Rg is allowed, with a flatter emissivity of q = 2.0+0.4 −0.7. Overall the fit statistic is good, with χ2/dof = 1983/1879 (Table 3, model 1); if the broad line is not included and the data are refit- ted then the fit is considerably worse (χ2/dof = 2027/1882, Table 3, model 2). Thus the broad line is required in the data at > 99.99% confidence. We also note that a simple broad Gaussian profile produces a very slightly worse fit than a diskline profile (χ2/dof = 1987/1879). In order to obtain tighter constraints on the diskline, a combined fit was then performed with the Suzaku and XMM-Newton EPIC-pn data. For this fit the inner radius of the disk is 37+25 −10Rg (for a fixed emissivity of q = 3), the inclina- tion angle is 53+7 ◦, while the line EW is 60 ± 11 eV. A −9 summary of the line parameters is shown in Table 2. Finally it has been previously suggested from XMM- Newton studies of the iron line profile in AGN that com- plex absorption (either ionized or partially covering the X-ray source) could provide an alternative explanation for the relativistic iron line (Reeves et al. 2004; Pounds et al. 2004; Turner et al. 2005). In order to test this hy- pothesis we fitted an iron K-shell bound -- free absorption edge constrained between 7.1 -- 9.3 keV (the known energy range from Fe i up to Fe xxvi), in order to mimic the additional opacity of an ionized absorber in the iron K bandpass. While it was possible to fit a weak iron K- shell edge to the Suzaku data (at 7.2 ± 0.1 keV with an optical depth of τ < 0.05), the fit was considerably worse than for a diskline (χ2/dof = 2020/1880 with an edge vs. χ2/dof = 1983/1879 with a disk-line). Similarly we also attempted to fit the spectrum with a neutral partial cov- ering absorber (the pcfabs model in xspec), which also resulted in a worse fit (χ2/dof = 2020/1880) with an ab- sorption column of ∼ 5 ×1022 cm−2 and a covering fraction of ∼ 10%. Note that there is no evidence for ionized ab- sorption lines and/or edges from lighter elements below the iron K band, either from the Suzaku XIS spectrum, or from the XMM-Newton EPIC-pm and Chandra/HETG spectra (Braito et al. 2006). Thus other than the neu- tral line of sight column density towards MCG -5-23-16 (of 1.6 × 1022 cm−2) which has been included in the spec- tral model, there is no evidence of complex absorption 8 Reeves et al. [Vol. , the strength of the Compton hump. The iron abundance was found to be sub-Solar, with AFe = 0.40 ± 0.12 relative to the Wilms et al. (2000) value. We can rule out a solar- abundance of iron with a high degree of confidence; fixing the abundance to AFe = 1 (but still allowing R to vary) worsened the fit statistic by ∆χ2 = 28, as the Fe K edge in the reflection model at 7.1 keV is then too deep to model the data well. Figure 7 (top panel) shows the confidence contours of R against the iron abundance. 3.4.1. Consistency Checks In the reflection fit, the constant cross normalization factor of the HXD over the XIS FI spectrum was also allowed to vary while fitting the reflection component. Figure 7 (lower panel) shows the confidence contours be- tween the constant and the reflection fraction (R). We find that in the best-fit described above, there is a small con- stant offset of 1.12 ± 0.05 between the HXD and XIS. This is consistent with calibration observations, the XIS+PIN spectrum of the Crab (Kokubun et al. 2006) also shows a factor of 1.15 ± 0.01 offset for a HXD nominal pointing and 1.13 ± 0.01 for XIS nominal (with a typical systematic uncertainty of 0.02). As one would expect, if the offset between the HXD and XIS increases, then the amount of reflection required decreases. Nonetheless the contours show that even upon allowing for some uncertainty in the constant offset between HXD and XIS, the parameters of the reflection component are well constrained. Note that the reflection parameters (and errors) quoted throughout this paper account for the uncertainty between R and the constant offset, as is shown in Figure 7. In order to investigate the possible systematics of the HXD background on the reflection parameters, the same fits were also performed on the Suzaku data, using in- stead a second independent background model for the HXD/PIN, as was shown in Figure 2 (the blue data points). A reflection fraction of R = 1.3 ± 0.3 was found to be consistent with the above value within the statisti- cal errors, with the other spectral parameters unchanged. We also used the simultaneous RXTE/PCA spectrum as a mission-independent check on the strength of the reflec- tion component. The RXTE PCU 2 data were reduced with standard extraction criteria, identical to those used in Markowitz & Edelson (2004). In fitting the RXTE spec- trum, the cut-off energy was fixed at 200 keV and the iron abundance at 0.6× solar, as neither of these parameters can be constrained in the PCA data. The iron line param- eters were fixed at the Suzaku XIS values, while the nor- malization of RXTE with respect to Suzaku is allowed to vary. With RXTE the reflection fraction was constrained to R = 1.0+0.7 −0.4, while Γ = 2.05 ± 0.12, consistent with the Suzaku observation. Overall in the HXD band, the model extrapo- lated flux from 15-100 keV of MCG -5-23-16 is 1.9 × 10−10 erg cm−2 s−1. This is consistent with the 15-100 keV flux determined from the Swift BAT survey of 1.6 × 10−10 erg cm−2 s−1 (Markwardt et al. 2005) and also with a flux of 1.8 × 10−10 erg cm−2 s−1 obtained from Integral (Bassani et al. 2006). The consistency of the flux mea- ratio residuals of Fig. 6. Data/model the spectrum of MCG -5-23-16, in the iron K-shell band. Panel (a) shows the residuals after fitting a power-law with a reflection component of R = 1.1. Panel (b) shows the residuals after the narrow Kα and Kβ Fe emission lines at 6.40 keV and 7.05 keV have been fitted. A broad excess is observed both red-wards and blue-wards of 6.4 keV. Panel (c) shows the residuals after the addition of a broad iron Kα line centered at 6.4 keV. The fit in the iron K band is now good, for completeness a weak Compton shoulder to the narrow Fe Kα line core has also been included. Note that the spectra have been re-binned by a factor of ×25 for clarity. that effects the detection of the broad iron K line. 3.4. Properties of the Reflection Component The neutral reflection component is also very well con- strained in the above model, with a best-fit value of R = 1.1 ± 0.2 with a lower-limit to the cut-off energy of Ecut > 170 keV. Note that if the GSO data-point between 50 -- 100 keV is included then Ecut = 270+150 −70 keV (the reflec- tion parameters are unchanged), which is consistent with the Beppo-SAX measurement (Perola et al. 2002; Balestra et al. 2004) for this source. Note the continuum photon index, cutoff energy, iron abundance, and a HXD/XIS constant normalization factor are allowed to vary indepen- dently along with the strength of the reflection hump (R). Interestingly the iron abundance of the reflector can also be constrained from the ratio of the neutral iron K edge to No. ] Revealing the High Energy Emission From MCG -5-23-16 9 O viii Ly α, O vii RRC, Ne ix (forbidden) and iron L-shell emission. The modeling of XMM-Newton RGS spectrum will be discussed in detail in a subsequent paper (Braito et al. 2006). However it seems plausible that the steep soft spectrum is the combination of a scattered power-law (with Γ ∼ 2 equal to the primary continuum) superim- posed on photoionized emission, similar to other Seyfert 2 s (Turner et al. 1997; Sambruna et al. 2001; Kinkhabwala et al. 2002; Pounds & Page 2005; Bianchi et al. 2005). Spectral fits to the XMM-Newton EPIC and RGS data confirm that this is likely to be the case. 4. Variability of the Iron Line and Reflection Component 4.1. Short-term Variability To test the variability of the iron line and reflection component over the timescale of the observation, the Suzaku data were split into high and low spectra for both XIS FI and HXD/PIN respectively. A threshold count rate in XIS 3 of > 3.7 counts s−1 and < 3.1 counts s−1 was used for the high and low flux spectra respectively, while spectra were then extracted with identical GTIs for XIS 0,2, HXD/PIN and the HXD/PIN background model. The XIS FI spectra were co-added as above. The resultant fluxes for the high and low spectra are then 9.9 ×10−11 erg cm−2 s−1 and 7.1 ×10−11 erg cm−2 s−1 respectively in the 2 -- 10 keV band; in the 15-100 keV band the fluxes are 2.0 × 10−10 erg cm−2 s−1 and 1.4 × 10−10 erg cm−2 s−1. The data were fitted with the same baseline model as above. The best fit parameters for the low and high flux spectra are shown in Table 2. In particular the in- put continuum normalization of the reflection compo- nent was tied at the mean value for the observation (3.2 × 10−2 photons cm−2 s−1 keV−1 at 1 keV). Then if the reflection component varies with the continuum, an in- crease in the relative R value (by about 40%) with respect to the mean continuum normalization should be observed in the high flux spectrum, compared to the low spectrum. However the resulting reflection component appears to be constant for the high and low state spectra, with values of R = 1.2 ± 0.2 and R = 1.3 ± 0.2 respectively. Similarly the iron line fluxes of the broad and narrow components are consistent with being constant, although some variability cannot be excluded. As a further check, the difference spectrum of the high minus low flux spectra was extracted, using both the XIS and HXD/PIN. The difference spectrum shows the variable component of the emission from MCG -5-23-16 modified by absorption, with the constant components in the spectrum being subtracted. The resulting differ- ence spectrum was fitted extremely well with a simple absorbed power-law of photon index Γ = 1.85 ± 0.09 and NH = 1.7 ± 0.1 × 1022 cm−2, while the fit statistic is ex- cellent (χ2/dof = 97.3/156). Within the errors, this is consistent with the continuum parameters derived from the mean spectrum. The difference spectrum is plotted in Figure 8. There are no residuals present in the iron K band Fig. 7. Contour plots showing the 68%, 90%, 99% and 99.9% confidence levels of the reflection fraction (R). In the top panel we show R versus the iron abundance (compared to the abundances in Wilms et al. 2000), demonstrating that the iron abundance is sub-Solar. The lower panel plots R ver- sus the constant normalization factor between the HXD/PIN and XIS FI, which shows that the constant factor is close to 1.1 and does not strongly effect the strength of the reflection component. surements between missions is a further indication of the robustness of the Suzaku HXD data. 3.5. The Soft X-ray Spectrum The XIS 1 were added to the spectral fit to provide better constraints on the soft spectrum. The soft un- absorbed power-law component below 1 keV is found to be steep with Γ = 3.1 ± 0.3. Furthermore the simultane- ous XMM-Newton EPIC-pn and MOS data also show a soft excess, with a photon index of Γ ∼ 3. In addition the XIS 1 data require an emission line at 0.92 ± 0.01 keV, with a flux of 1.5 ± 0.2 × 10−5 photons cm−2 s−1. At this energy the line may originate from Ne ix, but could also be blended with Fe L-shell emission. There are not suf- ficient counts to determine if there is O K-shell emis- sion in the XIS. However the long (100 ks) XMM-Newton RGS exposure simultaneous with the Suzaku observation shows a spectrum dominated by narrow emission lines be- low 1 keV, with lines from N vii Ly α, O vii (forbidden), 10 Reeves et al. [Vol. , respect to Solar was fixed at 0.6, as found earlier. In this case the reflection parameters appear consistent with each other, with R = 1.0 ± 0.2 and R = 0.7 ± 0.2 for Suzaku and Beppo-SAX respectively. The underlying photon index is Γ = 1.89 ± 0.04, with a high energy cut-off of 200+80 −50 keV. The flux levels of the observations are very similar, the 2- 10 keV fluxes being 8.8 × 10−11 erg cm−2 s−1 (Suzaku) vs 9.2 × 10−11 erg cm−2 s−1 (Beppo-SAX). One would expect consistent reflection values, given that the object was ob- served at similar fluxes. The iron line parameters measured in this paper also appear to be compatible with the values measured from previous observations; e.g. see Dewangan et al. (2003); Balestra et al. (2004) for comparison. The narrow iron Kα line flux is in excellent agreement with the previous values (between 4 − 8 × 10−5 photons cm−2 s−1), while the EW of the broad line measured from Suzaku (∼ 60 eV) is also consistent with the previous shorter XMM-Newton obser- vations within the errors (Dewangan et al. 2003; Balestra et al. 2004). This apparent lack of long term variability is perhaps not surprising considering this AGN has remained at a similar flux level (typically 7 −10 ×10−11 erg cm−2 s−1 in the 2-10 keV band). The source was only at a low flux level during the 1980s (see Figure 1, Mattson & Weaver 2004), as observed during the Ginga observations (Nandra & Pounds 1994) when the 2-10 keV flux reached a low of 2 × 10−11 erg cm−2 s−1. 5. Modeling the Reflection Component The Suzaku spectrum clearly shows two velocity width components to the iron line profile; the narrow core being unresolved with Suzaku and even Chandra/HETG, while the broad component has a velocity width of ∼0.1c. Thus it is possible that both components contribute towards the reflection spectrum, with the narrow line originating from distant matter (e.g. the putative torus) and the broad iron line from the accretion disk. Indeed one might expect an accretion disk reflector to contribute to the reflection spectrum, given the robust detection of the broad iron line. This hypothesis was tested by modeling the 0.4 -- 50 keV Suzaku spectrum with a dual reflection model; with an ionized disk reflector (using the reflion model, Ross & Fabian 2005) and a neutral distant reflector (using the pexrav model, Magdziarz & Zdziarski 1995). The narrow iron Kα, Kβ lines as well as the Compton shoulder were added to the model as described previously. The broad iron line is included as part of the ionized disk model. The ionized disk reflector is blurred by a Kerr metric from around a maximally rotating black hole (Laor 1991). A disk emissivity of q = 3 is assumed for the disk reflector, with an outer radius fixed at 400Rg. The best fit param- eters for the disk reflector are then R = 0.95+0.4 −0.5 for the reflection fraction, θ = 42 ±4◦ for the disk inclination angle with an inner radius truncated at Rin = 20+45 −11 Rg. The un- derlying continuum is Γ = 1.98 ± 0.04, consistent with the previous fits. The ionization parameter is consistent with neutral or low ionization iron (ξ < 60 erg cm s−1). Note Fig. 8. The difference spectrum between high and low flux states of MCG -5-23-16 in the Suzaku observation. This can be well represented by an absorbed power-law, with no iron K emission or reflection. Thus the power-law continuum is variable, while there appears to be no variable component of the reflection spectrum. between 6 -- 7 keV or any excess counts in the HXD/PIN difference spectrum above 10 keV, consistent with their being no variable component to the iron line or the reflec- tion component. From the difference spectrum, we can derive an upper-limit to any variable portion of the reflec- tion spectrum of R < 0.25, or < 20% of the total reflection component. The amplitude of the reflection variability is thus smaller than the continuum variability. Note that there is no excess of counts below 1 keV in the difference spectrum, also consistent with the soft excess being con- stant. 4.2. Long-Term Variability of the Reflection Component The long-term variations in the reflection component was investigated by comparing the Suzaku observation with the Beppo-SAX observation in 1998 (Perola et al. 2002; Balestra et al. 2004). The value of the reflection fraction in MCG -5-23-16 reported by Perola et al. (2002) is R = 0.66+0.25 −0.20, for an inclination of cosθ = 0.9. When converted to the inclination of 45◦ adopted in this paper, then their reflection value is R = 0.85+0.32 −0.25, which within the errors is consistent with the R value reported here from the Suzaku observation. We note that Balestra et al. (2004) find a lower R value of 0.45 (for a similar in- clination of 42◦) for the same Beppo-SAX observation. Note both of these papers assume solar abundances for the reflection component. We also performed a combined fit between the 1998 Beppo-SAX observation and the 2005 Suzaku observation. The continuum parameters (e.g. photon index and cut- off energy) as well as the iron line parameters were tied between the Beppo-SAX and the Suzaku datasets, allow- ing only the relative continuum fluxes to vary. The in- put continuum normalization to the reflection component was taken as the average of the Suzaku and Beppo-SAX values. The cross normalization between SAX/PDS and SAX/MECS was fixed at 0.87. The iron abundance with No. ] Revealing the High Energy Emission From MCG -5-23-16 11 spectrum is just fitted with the distant reflector includ- ing the narrow iron line core, then the fit statistic is also worse (χ2/dof = 2027/1882, model 2, Table 3), while the strength of the distant reflector increases to R = 1.5 ± 0.2 in order to compensate for the lack of the disk reflector and broad iron line. 6. Discussion and Conclusions The long Suzaku observation of the bright, Seyfert galaxy MCG -5-23-16 has revealed all the emission com- ponents expected in the X-ray spectrum of an obscured Seyfert galaxy in the context of AGN unified schemes (Antonucci 1993). A canonical high energy power-law (Γ = 1.9 − 2.0) continuum is observed, absorbed at low energies by Compton-thin matter and rolling over at high energies with a cut-off energy of > 170 keV. At soft X-ray energies a steep unabsorbed soft excess is ob- served, probably a combination of scattered power-law emission and photoionized line emission from distant gas. Such a component is common in many Seyfert 2 galaxies (Turner et al. 1997; Sambruna et al. 2001; Kinkhabwala et al. 2002; Pounds & Page 2005; Bianchi et al. 2005). Importantly, the excellent coverage of Suzaku at high en- ergies has allowed us to obtain extremely accurate con- straints on the iron K line profile and the high energy reflection component. Without a direct measurement of the reflection hump above 10 keV, the amount of reflec- tion present in the spectrum would be uncertain, making the determination of both the underlying continuum and the iron K line parameters somewhat degenerate. Due to the broad bandpass and high signal to noise of the Suzaku data, both the broad and narrow components of the iron Kα line as well as the reflection hump are unambiguously detected. The broad line is clearly resolved, with a veloc- ity width of σv = 20000 km s−1 (or 46000 km s−1 FWHM), which requires emission within ∼ 100Rg of the putative massive black hole in MCG -5-23-16. 6.1. The Geometry and Location of the Reprocessing Matter in MCG -5-23-16 The high signal to noise of the Suzaku XIS spec- trum, together with the simultaneous Chandra/HETG observation, limits the width of the narrow iron Kα line core. The upper limit obtained with Suzaku is σv < 2200 km s−1. For a black hole mass of 5 × 107M⊙ (Wandel & Mushotzky 1986), this corresponds to a distance of > 0.05 pc or > 2 × 104Rg. At this distance and given the detection of the distant reflection component observed in the HXD, a likely origin of the line is from fluorescence and scattering off material in a Compton-thick molec- ular torus (Ghisellini et al. 1994; Krolik et al. 1994). Alternatives to the torus could be a bi-conical outflow (Elvis 2000), or reflection off the outer radii of a warped disk (Nayakshin 2005). Two zones of Compton thick matter may exist in MCG - 5-23-16, the accretion disk being responsible for the broad iron Kα line, while the torus is the likely candidate for the distant reprocessor. This was first suggested for this Fig. 9. Best fit dual reflection model to the Suzaku spectrum of MCG -5-23-16. The solid red line shows the total emission from the model, the dashed green line shows the primary di- rect cut-off power-law continuum, while at low energies, the dashed line represents the scattered power-law. The relativis- tically blurred ionized disk reflector is plotted with the solid blue line, while the distant reflector is shown as a dashed blue line. The narrow iron K emission lines (Kα, Kβ and the Compton shoulder) are shown as solid black lines. The dual reflector model has reflection fractions of R ∼ 0.9 and R ∼ 0.5 for the disk and distant components respectively. The data are consistent with equal contributions from the disk and dis- tant reflectors, within the errors. that if we allow the disk emissivity to vary, then the best fit value is q = 2.3+0.9 −0.5 and subsequently an inner disk ra- dius extending down to 6Rg (the last stable orbit around a Schwarzschild black hole) or even 1.2Rg (for a maximal Kerr black hole) cannot be ruled out. The iron abundance was assumed to be the same value for both the disk and distant reflection components; the best fit value found was AFe = 0.65 ± 0.15. The neutral distant reflection component has a reflection fraction R = 0.5 ±0.3 (assuming a 45◦ inclination), while the equivalent width of the narrow iron Kα line is 77 ± 11 eV. For a plane parallel slab, illuminated by a continuum of Γ = 2 and viewed at 45◦ incidence, the predicted iron line equivalent width for a neutral reflector varies between 60 -- 70 eV (for 0.5× solar abundance), to 110 eV (for solar abundance), consistent with the measured value for the narrow iron line (Matt 2002). The overall dual reflection model is shown in Figure 9, while the fit statistic of this model fitted to the Suzaku spectrum is good (χ2/dof = 1980/1878, model 3, Table 3.). Note that a constant normalization offset factor between HXD/PIN and XIS was also included as a fit parameter, this was found to be 1.11 ± 0.04, consistent with what was found in the previous section. Both reflection components are required to model the Suzaku data. If the distant reflector and narrow iron line are removed and the spectrum refitted with just the ionized disk emission, then the fit statistic is much worse (χ2/dof = 2261/1882, model 4, Table 3), while the disk reflection fraction becomes rather large (R = 1.9). Alternatively if the disk reflector is removed (which in- cludes the emission from the broad iron line) and the 12 Reeves et al. [Vol. , Table 3. Comparison of iron line and reflection model fits to the mean spectrum. See text for model details. Model Description 1. 2. 3. 4. Neutral Reflection (pexrav) + Narrow and Broad Iron lines Neutral Reflection + Narrow Iron Line only Ionized Disk Reflection + Neutral reflection and Narrow Iron line Ionized Disk Reflection only χ2/dof 1983/1879 2027/1882 1980/1878 2261/1882 source by Weaver et al. (1997) on the basis of a two com- ponent model fit to the ASCA iron line profile. The dual reflection fits presented above suggest that the torus is likely Compton-thick (i.e. NH > 1024 cm−2), while the iron abundance in the torus is about half the solar value. The likely geometry for MCG -5-23-16 is that we are viewing through Compton-thin material at the edge of the torus which is Compton-thick at angles along the plane of the accretion disk. This picture seems consistent with the inclination angle derived from the diskline and disk reflec- tion fits to the broad iron line, of about 45◦. Similarly Weaver & Reynolds (1998) also found inclination angles in the range 40 -- 50◦ for the iron line fits to a sample of 4 Compton-thin Seyfert 2s (including MCG -5-23-16). Indeed the fact that MCG -5-23-16 displays broad Paschen β in the infra-red (Goodrich 1994) is consistent with view- ing the inner BLR through a moderate line of sight col- umn. An interesting question is how many of the Compton- thin Seyfert 2 galaxies show evidence for a Compton-thick reprocessor? The most detailed study to date of the high energy emission from Compton-thin Seyfert 2s was in an analysis of Beppo-SAX observations by Risaliti (2002). In the large majority of cases in this sample (17 out of 21 objects) a significant detection of Compton reflection has been found, while the authors favor a distant origin for the reflector from its apparent lack of variability. On the other hand the line of sight column densities of the Compton-thin Seyferts is too low (e.g. NH < 1023 cm−2) to have sufficient efficiency for Compton down -- scattering high energy photons. Thus the scenario whereby at least two zones of absorbing matter exists, with the Compton- thick zone out of the direct line of sight, would appear to be fairly commonplace in Compton-thin AGN. In the case of MCG -5-23-16, the column density of the line of sight absorber does not appear to significantly change with time; the column density measured by Suzaku (1.65 ± 0.03 × 1022 cm−2) lies close to the mean value of ∼ 1.7 ×1022 cm−2 observed by ASCA, Beppo-SAX, XMM- Newton and Chandra from 1994 -- 2001, with little scatter (Balestra et al. 2004) and it also agrees well with earlier measurements (Risaliti et al. 2002). Thus this absorber is likely to reside at large distances from the black hole (e.g. the torus or even the host galaxy) and is likely not to be clumpy. Finally we note that observations of Seyfert 2s, where the AGN appears to change from Compton- thick (i.e. reflection dominated) to Compton-thin or vice versa, also argues for the existence of at least two zones of distant absorbing matter (Matt et al. 2003). While the Compton-thick reprocessor could originate from par- sec scale material such as the torus, it is possible that the Compton-thin X-ray absorber originates from circumnu- clear gas or dust on a larger (e.g. ∼ 100 pc) scale within the AGN host galaxy (Maiolino et al. 1995; Malkan et al. 1998; Matt 2000; Guainazzi et al. 2005). 6.2. The Nature of the Iron K Line Emission The lack of variability of the reflection spectrum in MCG -5-23-16 on short timescales is also consistent with part of the reflection originating in distant matter. As the high -- low difference spectrum illustrates, the only variable part of the spectrum appears to be the intrin- sic power-law from the disk/corona. While the distant reprocessor is not expected to vary rapidly, the lack of short-term variability of the broad line in MCG -5-23-16 is curious. The best studied example to date is MCG - 6-30-15, where the relativistic iron line also appears not to vary despite strong continuum fluctuations (Vaughan & Fabian 2004), with gravitational light bending of the continuum photons near the black hole event horizon be- ing one possible explanation (Miniutti & Fabian 2004). In MCG -5-23-16 the iron line emission is much less cen- trally concentrated, with the line profile characterized by an inner radius of about ∼ 20 − 40Rg, unlike for MCG - 6-30-15 where the disk emission is likely to extend into 6Rg or even 1.2Rg (Wilms et al. 2001; Fabian et al. 2002). Taking a black hole mass estimate of 5 × 107M⊙ and if the line emission originates from radii > 20Rg, then the rever- beration timescale for the broad iron line will be > 104 s, which should be observable over the Suzaku observation (of 220 ks duration). One possibility is that the contin- uum flux variations are not strong enough (about 40%) in this source to produce a detectable difference in the broad iron K line flux. Suzaku observations of other bright, vari- able Seyfert galaxies with confirmed broad iron lines will be able to address this issue. An early indication is given by the deep (350 ks) Suzaku observation of MCG -6-30-15, which despite the rapid continuum flux variability, shows no strong variations in the iron line or reflection hump with continuum flux (Miniutti et al. 2006). Indeed the inner radius implied from the iron line pro- file could suggest that the inner disk may be missing or truncated in MCG -5-23-16. If the accretion rate is low enough, the innermost disk can transition to an ad- vective dominated accretion flow or ADAF (Narayan & Yi 1995). In MCG -5-23-16 the 2 -- 10 keV X-ray luminos- ity is 1.5 × 1043 erg s−1. Thus if the 2 -- 10 keV represents ∼ 5% of the bolometric output in a typical AGN (Elvis et al. 1994), then the bolometric luminosity of MCG -5-23-16 is of the order ∼ 3 × 1044 erg s−1. For a black hole mass of No. ] Revealing the High Energy Emission From MCG -5-23-16 13 5 × 107M⊙, the accretion rate of MCG -5-23-16 is ∼ 5% of the Eddington rate. Interestingly, this may be close to the transition rate between high/soft and low/hard states in Galactic black hole sources, thought to be a few percent of Eddington (Maccarone 2003). So it may be plausible that the inner disk could be truncated in MCG -5-23-16 at a few tens of gravitational radii. Conversely it is perhaps thought that most X-ray bright Seyfert galaxies are more analogous to high state black hole sources, at least based on their power density spectra (Uttley & McHardy 2005). However an optically-thick disk may extend inwards to the last stable orbit in MCG -5-23-16. If the emissivity of the X-ray source illuminating the disk is fairly flat (e.g. varying as R−2), then the disk may extend all the way to an inner radius of 6Rg or even 1.2Rg. In the spectral fits, this provides an equally acceptable fit to the data. A flat emissivity profile could occur if the illuminating source is located high above the disk in the "lamp-post" geometry, or if the X-ray emission is dissipated from flares further out on the disk. Furthermore it is possible that matter in the innermost disk radii is close to fully ionized. The reflection spectrum of a highly ionized disk or slab can be largely featureless and resemble that of the input contin- uum, especially if iron becomes fully ionized down to sev- eral Thomson depths (Nayakshin et al. 2000; Ballantyne et al. 2001). In this case the inner disk radius could corre- spond to a characteristic radius at which the disk surface becomes fully ionized. An alternative possibility is that the broad iron line may not even originate from the accretion disk. The emission could originate from reprocessing in either a spherical dis- tribution of clouds (Guilbert & Rees 1988) or a bi-conical outflow (Sulentic et al. 1998; Elvis 2000). Ultra-fast out- flows have been claimed in a handful of AGN, on the basis of blue-shifted high ionization iron K absorption lines or edges; e.g. PG 1211+143 (Pounds et al. 2003b), PDS 456 (Reeves et al. 2003), IC 4329a (Markowitz et al. 2006). Indeed the long December 2005 XMM-Newton observa- tion of MCG -5-23-16 shows evidence for a variable, blue- shifted iron Kα absorption line at 7.8 keV (Braito et al. 2006), which could plausibly arise from such an outflow. A broader issue is the apparent lack of broad iron lines detected in recent XMM-Newton observations of bright Seyfert galaxies. Despite the high quality of the obser- vations, sometimes only a narrow line is present (Pounds et al. 2003a; Schurch et al. 2003; Bianchi et al. 2004). In some cases broad residuals are present in the iron K band, but the continuum curvature due to a high column warm absorber and a red-wing of a relativistic iron line are difficult to distinguish (Reeves et al. 2004; Pounds et al. 2004; Turner et al. 2005). In the example of MCG - 5-23-16, with Suzaku it is clear that a broad iron line is present. However in a much shorter observation of a weaker source such a line may not be detected, or its modeling could be ambiguous where there is no bandpass above 10 keV. As discussed above, the disk lines may be weaker than expected if the inner accretion disk is close to fully ionized; this will be dependent both on the il- luminating continuum and the X-ray emission geometry (Nayakshin et al. 2000). Also if the iron abundance is less than Solar (as measured in MCG -5-23-16), the broad line will be weaker. In AGN X-ray spectra where complex or even partial covering absorption appears to be present (e.g. NGC 3516 Turner et al. 2005, NGC 4051 Pounds et al. 2004), one of the major problems in assessing the contribution of the broad iron line is in the uncertainty in determining the underlying continuum. Suzaku presents the best capa- bilities of the current X-ray missions for resolving this issue, by revealing the true shape and level of the under- lying continuum and reflection component through broad- band observations with coverage above 10 keV, while also achieving high signal to noise (at least equal to XMM- Newton) in the iron K band. Encouragingly there appears to be a large number of hard X-ray selected AGN emerg- ing from the Swift/BAT All Sky Survey (Markwardt et al. 2005), which could eventually number as a many as ∼ 200 Type I and Type 2 AGN at a limiting flux level of ∼ 10−11 erg cm−2 s−1. Most of these AGN will be bright enough for detailed study both in the iron K bandpass and above 10 keV with Suzaku. These sources can form the basis of our future understanding of the iron line and Compton reflection associated with the innermost regions of AGN and for testing our understanding of Unified mod- els and AGN evolution. References Antonucci, A. 1993, ARA&A, 31, 473 Ballantyne, D.R., Ross, R.R., & Fabian, A.C. 2001, MNRAS, 327, 10 Balucinska-Church, M. & McCammon, D. 1992, ApJ, 400, 699 Bassani, L. et al. 2006, ApJ, 636, L65 Balestra, I., Bianchi, S., & Matt, G. 2004, A&A, 415, 437 Bianchi, S., Matt, G., Balestra, I., Guainazzi, M., & Perola, G.C. 2004, A&A, 422, 65 Bianchi, S., Miniutti, G., Fabian, A. C., & Iwasawa, K. 2005, MNRAS, 360, 380 Boldt, E. 1987, Phys. Rep., 146, 215 Churazov, E. et al., 2006, A&A, in press Dewangan, G.C., Griffiths, R.E., & Schurch, N.J. 2003, ApJ, 592, 52 Dickey, J.M., & Lockman, F.J. 1990, ARA&A, 28, 215 Elvis, M., et al., 1994, ApJS, 95, 1 Elvis, M. ApJ, 2000, 545, 63 Fabian, A.C., Rees, M.J., Stella, L., & White, N.E. 1989, MNRAS, 238, 729 Fabian, A.C., et al., 2002, MNRAS, 335, L1 George, I.M., & Fabian, A.C. 1991, MNRAS, 249, 352 Ghisellini, G., Haardt, F., & Matt, G. 1994, MNRAS, 267, 743 Goodrich, R.W., Veilleux, S., & Hill, G.J. 1994, 422, 521 Gruber, D.E., Matteson, J.L., Peterson, L.E., & Jung, G.V. 1999, ApJ, 520, 124 Guainazzi, M., Matt, G., Perola, G.C. 2005, A&A, 444, 119 Guilbert, P.W. & Rees, M.J. 1988, MNRAS, 233, 475 Kinkhabwala, A. et al., 2002, ApJ, 575, 732 Kokubun, M. et al., 2006, PASJ, this issue Koyama, K. et al., 2006, PASJ, this issue Krolik, J.H., Madau, P., & Zycki, P.T. 1994, ApJ, 266, 653 Kushino, A., Ishisaki, Y., Morita, U., Yamasaki, N.Y., Ishida, M., Ohashi, T., & Ueda, Y. 2002, PASJ, 54, 327 14 Reeves et al. [Vol. , Weaver, K.A., Yaqoob, T., Mushotzky, R.F., Nousek, J., Hayashi, I., & Koyama, K. 1997, ApJ, 474, 675 Wegner, G., et al., 2003, AJ, 126, 2268 Wilms, J., Allen, A., McCray, R. 2000, ApJ, 542, 914 Wilms, J., Reynolds, C.S., Begelman, M.C., Reeves, J., Molendi, S., Staubert, R., & Kendziorra, E. 2001, MNRAS, 328, L27 Yaqoob, T., & Padmanabhan, U. 2004, ApJ, 604, 63 Laor, A. 1991, ApJ, 376, L90 Lightman, A.P., & White, T.R. 1988, ApJ, 335, L57 Maccarone, T.J. 2003, A&A, 409, 697 Magdziarz, P., & Zdziarski, A. 1995, MNRAS, 273, 837 Maiolino, R., Ruiz, M., Rieke, G.H., Keller, L.D. 1995, ApJ, 446, 561 Malkan, M.A., Gorjian, V., Tam, R. 1998, ApJS, 117,25 Markowitz, A.. & Edelson, R. 2004, ApJ, 617, 939 Markowitz, A., Reeves, J.N., & Braito, V. 2006, ApJ, 646, 783 Markwardt, C.B., Tueller, J., Skinner, G.K., Gehrels, N., Barthelmy, S.D., & Mushotzky, R.F. 2005, ApJ, 633, L77 Matt, G. 2000, A&A, 355, 31L Matt, G. 2002, MNRAS, 337, 147 Matt, G., Guainazzi, M., Maiolino, R., 2003 MNRAS, 342, 422 Mattson, B.J., & Weaver, K.A. 2004, ApJ, 601, 771 Miniutti, G. & Fabian, A.C. 2004, MNRAS, 349, 1435 Miniutti, G. et al., 2006, PASJ, this issue Mushotzky, R. 1982, ApJ, 256, 92 Mitsuda, K. et al., 2006, PASJ, this issue Nandra, K., & Pounds, K.A. 1994, MNRAS, 268, 405 Nandra, K., George, I.M., Mushotzky, R.F., Turner, T.J., & Yaqoob, T. 1997, ApJ, 477, 602 Narayan, R. & Yi, I. 1995, ApJ, 452, 710 Nayakshin, S., Kazanas, D., & Kallman, T.R. 2000, ApJ, 537, 833 Nayakshin, S. 2005, MNRAS, 359, 545 Page, K.L., O'Brien, P.T., Reeves, J.N., & Turner, M.J.L. 2004, MNRAS, 347, 316 Perola, G.C., Matt, G., Cappi, M., Fiore, F., Guainazzi, M., Maraschi, L., Petrucci, P.O., & Piro, L. 2002, A&A, 389, 802 Pounds, K.A., Nandra, K., Stewart, G.C., George, I.M., & Fabian, A.C. 1990, Nature, 344, 132 Pounds, K.A., Reeves, J.N., Page, K.L., Edelson, R., Matt, G., & Perola, G.C. 2003a, MNRAS, 341, 953 Pounds, K.A., Reeves, J.N., King, A.R., Page, K.L., O'Brien, P.T., & Turner, M.J.L. 2003b, MNRAS, 345, 705 Pounds, K.A., Reeves, J.N., King, A.R., & Page, K.L. 2004, MNRAS, 350, 10 Pounds, K.A., & Page, K.L. 2005, MNRAS, 360, 1123 Reeves, J.N., O'Brien, P.T., & Ward, M. 2003, ApJ, 593, L65 Reeves, J.N., Nandra, K., George, I.M., Pounds, K.A., Turner, T.J., & Yaqoob, T. 2004, 602, 648 Reynolds, C.S. 1997, MNRAS, 286, 513 Risaliti, G. 2002, A&A, 386, 379 Risaliti, G., Elvis, M., & Nicastro, F. 2002, ApJ, 571, 234 Ross, R.R. & Fabian, A.C. 2005, MNRAS, 358, 211 Sambruna, R.M., et al., 2001, ApJ, 546, L13 Schurch, N., Warwick, R.S., Griffiths, R.S., & Sembay, S. 2003, MNRAS, 345, 423 Sulentic, J.W., Marziani, P., Zwitter, T., Calvani, M., & Dultzin-Hacyan, D. 1998, ApJ, 501, 54 Takahashi, T. et al., 2006, PASJ, this issue Tanaka, Y. et al. 1995, Nature, 375, 659 Turner, T.J., George, I.M., Nandra, K., & Mushotzky, R.F. 1997, ApJ, 488, 164 Turner, T.J., Kraemer, S.B., George, I.M., Reeves, J.N., & Bottorff, M C. 2005, 618, 155 Uttley, P. & McHardy, I.M. 2005, MNRAS, 363, 586 Vaughan, S. & Fabian, A.C. 2004, MNRAS, 348, 1415 Veron, P., Lindblad, P.O., Zuiderwijk, E.J., Veron, M.P., & Adam, G. 1980, A&A, 87, 245 Wandel, A., & Mushotzky, R.F. 1986, ApJ, 306, L61 Weaver, K.A., & Reynolds, C.S. 1998, ApJ, 503, L39
astro-ph/0605087
1
0605
2006-05-03T11:23:23
Probing the Pulsar Wind Nebula of PSR B0355+54
[ "astro-ph" ]
We present XMM-Newton and Chandra X-ray observations of the middle-aged radio pulsar PSR B0355+54. Our X-ray observations reveal emission not only from the pulsar itself, but also from a compact diffuse component extending ~50'' in the opposite direction to the pulsar's proper motion. There is also evidence for the presence of fainter diffuse emission extending ~5' from the point source. The compact diffuse feature is well-fitted with a power-law, the index of which is consistent with the values found for other pulsar wind nebulae. The morphology of the diffuse component is similar to the ram-pressure confined pulsar wind nebulae detected for other sources. The X-ray emission from the pulsar itself is described well by a thermal plus power-law fit, with the thermal emission most likely originating in a hot polar cap.
astro-ph
astro-ph
Draft version June 19, 2018 Preprint typeset using LATEX style emulateapj v. 6/22/04 6 0 0 2 y a M 3 1 v 7 8 0 5 0 6 0 / h p - o r t s a : v i X r a PROBING THE PULSAR WIND NEBULA OF PSR B0355+54 K.E. McGowan1,2, W.T. Vestrand3, J.A. Kennea4, S. Zane2, M. Cropper2, F.A. C´ordova5 Draft version June 19, 2018 ABSTRACT We present XMM-Newton and Chandra X-ray observations of the middle-aged radio pulsar PSR B0355+54. Our X-ray observations reveal emission not only from the pulsar itself, but also from a compact diffuse component extending ∼ 50′′ in the opposite direction to the pulsar's proper motion. There is also evidence for the presence of fainter diffuse emission extending ∼ 5′ from the point source. The compact diffuse feature is well-fitted with a power-law, the index of which is consis- tent with the values found for other pulsar wind nebulae. The morphology of the diffuse component is similar to the ram-pressure confined pulsar wind nebulae detected for other sources. The X-ray emission from the pulsar itself is described well by a thermal plus power-law fit, with the thermal emission most likely originating in a hot polar cap. Subject headings: pulsars: individual (PSR B0355+54) -- stars: neutron -- X-rays: stars 1. INTRODUCTION Isolated pulsars constitute one of the most power- laboratories for studying particle acceleration in ful A significant fraction of the energy astrophysics. from rotation-powered pulsars is converted into a wind (Rees & Gunn 1974), which travels at a velocity close to the speed of light. The interaction of this pulsar wind with the ambient medium produces a shock and acceler- ation of the relativistic particles at the shock generates synchrotron emission. This non-thermal diffuse emission manifests itself as a pulsar wind nebulae (PWN) or ple- rion at radio and X-ray energies (e.g. Rees & Gunn 1974; Gaensler 2001). Due to the short synchrotron lifetimes of high energy electrons, X-ray emission from a PWN directly traces the current energetics of the pulsar. The spectral and morphological characteristics of an X-ray PWN therefore reveal the structure and composition of the pulsar wind and the orientation of the pulsar's spin axis and/or velocity vector. The middle-aged 156 ms radio pulsar PSR B0355+54 is known to emit X-rays (Helfand 1983; Seward & Wang 1988; Slane 1994) and gamma-rays (Bhat et al. 1990). Helfand (1983) reported the first detection in X-rays of the source using data from Einstein, stating that emission extended 5′ However, Seward & Wang (1988) analyzed the Einstein data and concluded that while there was evidence for weak emis- sion 1.7′ from the source position, emission from the pul- sar itself was not detected. Nevertheless, they did not rule out the possibility that the emission could be associ- ated with a PWN. Slane (1994) detected PSR B0355+54 in a 20 ks ROSAT observation, but owing to the lack of counts it was not feasible to perform a spectral analysis. from the pulsar. 1 School of Physics and Astronomy, Southampton University, Southampton, UK 2 Mullard Space Science Laboratory, University College of Lon- don, UK 3 Los Alamos National Laboratory, Los Alamos, NM 87545 4 Pennsylvania State University, 525 Davey Laboratory, Univer- sity Park, PA 16802, USA 5 Chancellor's Office, University of California, Riverside, CA 92521 Electronic address: [email protected] The analysis of the ROSAT data also led to the detection of faint extended emission ∼ 1.6′ from the pulsar posi- tion, but Slane (1994) did not believe there was enough evidence to support a link between the source and the extended emission. In this paper, we report on XMM-Newton and Chan- dra observations of PSR B0355+54 which we use to in- vestigate the presence of diffuse emission that can be attributed to a PWN. 2. OBSERVATIONS AND DATA REDUCTION PSR B0355+54 was observed with XMM-Newton on 2002 February 10 for 29 ks. We used data from the European Photon Imaging Camera (EPIC) PN instru- ment (Struder et al. 2001) for the spatial, spectral and timing analysis. The PN was configured in small win- dow mode and the thin blocking filter was used. Data from the MOS1 instrument (Turner et al. 2001) was also used for the spatial analysis. The MOS1 was operated in full window mode with the medium filter. We reduced the EPIC data with the XMM-Newton Science Analysis System (SAS version 6.1.0). In order to maximize the signal-to-noise ratio for our XMM-Newton observation, we filtered the data to include only single, double, triple and quadruple photon events for the MOS1, and only single and double photon events for the PN. Data were filtered to exclude events that may be incorrect, for ex- ample those next to the edges of the CCDs and next to bad pixels. We only included photons with energies in the range 0.3 − 10 keV. PSR B0355+54 was observed for 66 ks on 2004 July 16 with the ACIS-S array on Chandra in the very faint timed exposure imaging mode. We performed standard data processing using CIAO version 3.2. The data were filtered to restrict the energy range to 0.3 − 10 keV and to exclude times of high background. 3. SPATIAL ANALYSIS Initial inspection of the images created from the EPIC- PN and EPIC-MOS1 data show relatively strong emis- sion at the pulsar position and evidence for extended emission near to PSR B0355+54 (see Figure 1, top panel). We generated a mosaic of the PN and MOS1 im- 2 McGowan et al. Fig. 1. -- X-ray detection of PSR B0355+54 and its diffuse emission. Top panel: Gray-scale plot of the 0.3 − 10 keV XMM- Newton PN image. Middle panel: Gray-scale plot of the 0.3 − 10 keV Chandra ACIS image. The boxes define the regions used to extract spectra for the diffuse emission. Bottom panel: The same image as the middle panel with the contribution from the X-ray point source removed. The image is smoothed with a Gaussian of width ∼ 2′′. The arrow shows the direction of the proper motion of the pulsar and has a length of 20′′. The circle marks the "south west" source (see Section 3). ages and measured the X-ray source positions using the SAS source detection tool EDETECT CHAIN. We com- pared the positions of the field stars in our observation with the positions from optical catalogs to determine an astrometric correction. This correction was applied to the X-ray coordinates of the pulsar, resulting in R. A. = 03h58m53.s73, decl. = +54◦13′12.′′12 (J2000), with an rms error of 0.′′78. This position lies 1.′′6 from the radio position. To confirm the presence and examine the extent of the diffuse emission in the XMM-Newton data we have com- pared the detected PN emission with that for a point source. We calculated the intensity for the pulsar by using bilinear interpolation at regularly spaced points along the direction of proper motion of PSR B0355+54 (Chatterjee et al. 2004). We compared this profile with the XMM-Newton point-spread function (PSF) for the PN at 1.5 keV, which we generated using the King profile parameters included in the XMM-Newton calibration file Fig. 2. -- X-ray emission from PSR B0355+54 as a function of distance from the point source along the direction of the proper motion of the pulsar (solid line), compared to the instrument PSF determined at 1.5 keV and the location of the pulsar (dashed line). Top panel: XMM-Newton PN. Bottom panel: Chandra ACIS. "XRT3 XPSF 0006.CCF.plt"6. In Figure 2 (top panel) we show the profiles for the pulsar and the PN PSF. The positions of the X-ray sources for the ACIS data were determined using the CIAO source detection tool WAVDETECT. The image does not contain enough sources with known counterparts to perform an astro- metric correction to the coordinates. A point source is detected at R. A. = 03h58m53.s70, decl. = +54◦13′13.′′87 (J2000), which is 0.′′14 away from the radio pulsar po- sition. This source is consistent with being the X-ray counterpart of the pulsar. The ACIS image also reveals a faint tail of emission in the opposite direction to the pulsar's proper motion (see Figure 1, middle and bottom panels). Again we determined the net counts from the source and diffuse emission at regularly spaced intervals along the direc- tion of the pulsar's proper motion. We generated a PSF for PSR B0355+54 using the Chandra PSF library evaluated at 1.5 keV and the location relevant to our source. The PSF was normalized to the total counts in PSR B0355+54. We calculated the net counts for the PSF in the same intervals as for PSR B0355+54. The source and PSF profiles are shown in Figure 2 (bottom 6 See http://xmm.vilspa.esa.es/docs/documents/CAL-SRN-0100-0-0.ps.gz for more information The Pulsar Wind Nebula of PSR B0355+54 3 In order to investigate the properties of the X-ray emis- sion from PSR B0355+54 and the compact (≤ 50′′) dif- fuse nebula, we compared the spectra extracted from different spatial regions. Our results from the spatial analysis suggest that the core of the pulsar emission lies within 5′′ of the pulsar's position. However, in the case of the XMM-Newton data, this size of aperture does not contain enough counts for a meaningful analysis. The pulsar spectrum has been extracted from the XMM-Newton observation using a circular region of ra- dius 30′′, centered on the pulsar's radio position. The background was extracted from a region of similar size offset from the pulsar position. The total counts con- tained in the source region is 1143 with an estimated 562 from background. The spectrum was regrouped by re- quiring at least 50 counts per spectral bin. We created a photon redistribution matrix (RMF) and ancillary region file (ARF) for the spectrum. The subsequent spectral fit- ting and analysis was performed using XSPEC, version 11.3.1. We modeled the spectrum in the 0.5 − 9.0 keV range. Initially we fitted the spectrum with single-component models including absorbed power-law, blackbody and magnetized, pure H atmospheric (Pavlov et al. 1995) models. The spectrum is best-fitted with a power-law with index Γ = 1.5+0.5 −0.3 and column density NH = (0.50+0.36 −0.20) × 1022 cm−2. This value for the power-law index is similar to the values found for other PWNe (e.g. Kaspi et al. 2005). The Galactic hydrogen column in the direction of PSR B0355+54 is NH = 0.88 × 1022 cm−2. The fit results in an unabsorbed 0.3 -- 10 keV energy flux of (2.3+1.0 −0.7) × 10−13 ergs cm−2 s−1. We also fitted the spec- trum with blackbody plus power-law and atmospheric plus power-law models, both modified by photoelectric absorption. The multi-component models give similar values for reduced χ2, however the temperatures implied by the fits are poorly constrained. It is likely that the presence of the pulsar wind nebula, and being unable to separate the pulsar core and diffuse emission, effects our ability to constrain the thermal component in the spec- tral fits of the XMM-Newton data. The results of the XMM-Newton spectral fitting are given in Table 1. The XMM-Newton spectrum with the best-fitting power-law model is shown in Figure 4. For the Chandra data we extracted a spectrum for the core of the pulsar emission from a circular region of radius 5′′ centered on the radio position. An annulus centered on the pulsar position was used to extract the background, with inner and outer radii of 6′′ and 10′′, re- spectively. We find a total of 244 counts contained in the source region, with 29 counts attributed to background. We created the RMF and ARF files using standard CIAO tools. Before fitting the spectrum we regrouped the data, requiring a minimum of 15 counts per spectral bin. We fitted the spectrum in the energy range 0.5 − 7.0 keV using the same models as for the XMM-Newton data. In the first instance we let the neutral hydrogen column density be a free parameter; however this led to unrea- sonably small values for NH . Subsequently we fixed the column density at the value found from the power-law fit to the XMM-Newton spectrum of PSR B0355+54. We find that the Chandra spectrum can also be character- ized by a power-law. The model has a power-law index Fig. 3. -- XMM-Newton MOS1 X-ray emission from PSR B0355+54 as a function of distance from the point source, orientated 20◦ further North from the direction of the pulsar's proper motion (solid line), compared to the mean background (dot- ted line). panel). The XMM-Newton PN image shows emission ∼ 45′′ south west of the pulsar which could be a source, how- ever, there is no corresponding emission at this position in the Chandra ACIS image (see Figure 1). We per- formed wavelet analysis on the Chandra data which con- firms there is no source detected. Comparison of the XMM-Newton field of PSR B0355+54 with the Digitized Sky Survey does not show any optical source at the po- sition of the south west emission. If the emission in the XMM-Newton data is real, it suggests that the diffuse emission could be varying over time. We smoothed the Chandra ACIS image with a Gaus- sian of width ∼ 2′′ (Figure 1, bottom panel), the result- ing image suggests that there are two regions of enhanced diffuse emission -- one near to the pulsar and the other ∼ 10′′ away. The intensity profiles for the XMM-Newton and Chandra data indicate that the core of the X-ray emission lies within 5′′ of the pulsar position. Both emis- sion profiles indicate that the diffuse emission extends out to ∼ 50′′, with the bulk of the emission lying within 20 − 30′′ of PSR B0355+54. The profiles also show ev- idence for a dip in the emission at ∼ 10′′ agreeing with the smoothed ACIS image. Detection of emission from PSR B0355+54 at distances of 1.′6 to 5′ from the source position i.e. on a larger scale than shown in Figure 1, have been reported by Helfand (1983), Seward & Wang (1988) and Slane (1994), see also Tepedelenlioglu & Ogelman (2005). Visual inspection of a smoothed version of the MOS1 image indicates that there is a region of enhanced emission extending a few arcminutes south of the pulsar, orientated ∼ 20◦ closer to North than the diffuse emission we have reported above. The intensity of the MOS1 emission in this direction was determined using bilinear interpolation at regularly spaced intervals from the source. We show in Figure 3 the distribution of counts as a function of distance from the source compared to the mean background. Our anal- ysis suggests that there is an excess of counts at 1′ and ∼ 3 − 5′ from the point source. 4. SPECTRAL ANALYSIS 4 McGowan et al. TABLE 1 Best-fit parameters for the XMM-Newton spectrum of PSR B0355+54 Model NH Γ (1022 cm−2) PL BB BB + PL NSA NSA + PL 0.50+0.36 −0.20 < 0.01 0.23+0.69 −0.23 0.13+0.23 −0.11 0.49+0.21 −0.21 1.5+0.5 −0.3 · · · 0.8+0.3 −0.3 · · · 1.5+1.0 −0.3 T /T ∞ ef f (106 K) · · · 11.26+1.74 −1.51 7.66+2313 −4.64 7.66+0.01 −2.83 0.64+6.83 −0.75 χ2 ν (dof) FX (erg cm−2 s−1) (2.3+1.0 (1.4+0.1 (2.0+1.0 (1.4+0.1 (2.4+1.1 −0.7) × 10−13 −1.1) × 10−13 −1.1) × 10−13 −1.3) × 10−13 −1.0) × 10−13 0.7 (16) 1.0 (16) 0.7 (14) 0.9 (16) 0.8 (14) Note. -- The last column is the unabsorbed flux in the 0.3 -- 10 keV range. In the case of the NSA model the mass and radius of the neutron star are fixed at MN S = 1.4M⊙ and RN S = 10 km, respectively, and the magnetic field of the neutron star is fixed at B = 1012 G. The errors quoted are the 90% uncertainties. the source and the mass of the neutron star were fixed at D = 1.04 kpc and MN S = 1.4M⊙, respectively. The magnetic field of the neutron star was fixed at B = 1012 G (this is a good approximation since the pulsar mag- netic field as inferred from radio timing properties is B = 8.4 × 1011 G, Hobbs et al. 2004; Manchester et al. 2005). The unabsorbed 0.3 -- 10 keV energy flux for this fit is (1.5+54.0 −0.7 ) × 10−13 ergs cm−2 s−1. The Chandra spec- trum of the core emission from PSR B0355+54 is shown in Figure 5 with the best-fitting blackbody plus power- law model (top) and magnetized, pure H atmospheric plus power-law model (bottom). To analyze the compact diffuse emission we created a new events file in which the emission from the pulsar core was removed. We extracted a spectrum for the diffuse component from a rectangular region of 40′′ × 55′′, cen- tered on the emission and orientated along the direction of the pulsar's proper motion. The background was ex- tracted from a region of similar size offset from the diffuse emission. The diffuse component extraction region con- tains 1207 counts, with an estimated 414 counts due to background. We created the RMF and ARF files using standard CIAO tools. Before fitting the spectrum we re- grouped the data, requiring a minimum of 15 counts per spectral bin. We modeled the spectrum over 0.5 − 7.0 keV with an absorbed power-law, keeping the column density fixed at NH = 0.50 × 1022 cm−2. The best-fit has a power-law index of Γ = 1.4 ± 0.3 and unabsorbed 0.3 -- 10 keV energy flux of (1.7+0.8 −0.5) × 10−13 ergs cm−2 s−1. In order to investigate the possibility of spectral evolu- tion along the extended X-ray emission we created spec- tra for three regions of the compact diffuse emission. The sizes of the regions were chosen with the aim of having similar numbers of counts in each region. The three ex- traction regions, orientated along the direction of proper motion, are as follows, region 1: 40′′ × 18′′, contains a total of 396 counts with 130 attributed to background, region 2: 40′′ × 9′′, contains a total of 356 counts with 69 attributed to background, region 3: 40′′ × 28′′, con- tains a total of 454 counts with 214 attributed to back- ground (see Figure 1 (middle panel)). The background was extracted from the same region as above. We cre- ated response files for each region and regrouped the spectra, requiring a minimum of 15 counts per spectral Fig. 4. -- The XMM-Newton spectrum of PSR B0355+54 with best-fit power-law model. Also shown are the residuals from com- parison of the data to the model. of Γ = 1.9+0.4 −0.3 which is consistent within the 90% uncer- tainties to the value found from the XMM-Newton data. However, in the case of the Chandra data we find that a thermal plus power-law model provides a better fit statis- tically. The data are equally well-fitted by a blackbody plus power-law and a magnetized, pure H atmospheric (Pavlov et al. 1995, "nsa" model in XSPEC) plus power- law model. The results of the Chandra spectral fitting are given in Table 2. For the blackbody plus power-law model the best- fit parameters are a power-law index of Γ = 1.0+0.2 −1.0 and temperature of T = (2.32+1.16 −0.81) × 106 K. Us- ing a distance to the source of D = 1.04+0.21 −0.16 kpc (Chatterjee et al. 2004), this implies a blackbody emit- ting radius of 0.12+0.16 −0.07 km. This value is too small to be reconciled with the radius of the neutron star and would indicate that the origin of the emission is a hot polar cap. We find an unabsorbed 0.3 -- 10 keV energy flux of (6.4+19.3 −1.1 ) × 10−14 ergs cm−2 s−1. The magnetized, pure H atmospheric plus power-law model best-fit parameters are a power-law index of Γ = 1.5+0.5 −0.4, temperature of ef f = (0.45+0.20 −0.22) × 106 K, and a radius for the neutron T ∞ star of RN S = 7.2+7.2 −2.2 km. For this fit the distance to The Pulsar Wind Nebula of PSR B0355+54 5 Best-fit parameters for the Chandra X-ray emission of PSR B0355+54 TABLE 2 Region Model Γ Core 1.9+0.4 −0.3 · · · · · · 1.0+0.2 −1.0 PL BB BB + PL NSA NSA + PL 1.5+0.5 −0.4 1.4+0.3 −0.3 1.4+0.4 −0.4 1.5+0.3 −0.3 1.2+0.5 −0.4 Diffuse -- all PL PL Diffuse -- 1 PL Diffuse -- 2 Diffuse -- 3 PL T /T ∞ ef f (106 K) · · · 6.73+2.32 −1.87 2.32+1.16 −0.81 0.48 0.45+0.20 −0.22 · · · · · · · · · · · · RBB /RN S χ2 ν (dof) FX km · · · 0.01+0.03 −0.01 0.12+0.16 −0.07 9.5 7.2+7.2 −2.2 · · · · · · · · · · · · (erg cm−2 s−1) (4.9+1.6 (2.4+0.1 (6.4+19.3 (1.8+0.3 (1.5+54.0 (1.7+0.8 (5.7+3.3 (7.6+3.4 (5.3+4.4 −0.7) × 10−14 −0.1) × 10−14 −1.1 ) × 10−14 −0.5) × 10−13 −0.7 ) × 10−13 −0.5) × 10−13 −1.8) × 10−14 −1.9) × 10−14 −2.0) × 10−14 0.5 (34) 1.4 (34) 0.3 (32) 3.6 (34) 0.4 (32) 1.0 (50) 1.0 (16) 1.4 (16) 1.2 (17) Note. -- The neutral hydrogen column density has been fixed at NH = 0.50 × 1022 cm−2 in all of the fits. The last column is the unabsorbed flux in the 0.3 -- 10 keV range. In the case of the NSA model the distance to the source is fixed at D = 1.04 kpc, the mass of the neutron star is fixed at MN S = 1.4M⊙, and the magnetic field of the neutron star is fixed at B = 1012 G. The errors quoted are the 90% uncertainties. bin. The spectra were fitted in the 0.5 − 7.0 keV en- ergy range with a power-law and fixed column density of NH = 0.50 × 1022 cm−2. We find the best-fit power- law indices for regions 1 -- 3 are Γ = 1.4 ± 0.4, 1.5 ± 0.3 and 1.2+0.5 −0.4, respectively. Due to the uncertainties on the indices the presence of any spectral variability remains unclear. The unabsorbed 0.3 -- 10 keV energy fluxes are (5.7+3.3 −2.0) × 10−14 ergs cm−2 s−1, respectively. The Chandra spectra of the diffuse emission from regions 1 -- 3 are shown in Figure 6 (first -- third panel) with the best-fitting power-law mod- els. −1.9) × 10−14 and (5.3+4.4 −1.8) × 10−14, (7.6+3.4 We have also tried to determine if there are any spec- tral changes by using the hardness ratio h2.0, which is defined as the ratio of counts above 2.0 keV to that be- low 2.0 keV. For the whole of the compact diffuse com- ponent we find h2.0 = 0.97 ± 0.09. Regions 1, 2 and 3 have h2.0 = 1.05 ± 0.17, 0.91 ± 0.13 and 0.97 ± 0.18, re- spectively. Due to the uncertainties, no particular trends in hardness ratio can be determined from one region to the next. 5. TIMING ANALYSIS We barycentrically corrected the photon arrival times in the XMM-Newton PN event file before performing the temporal analysis. We extracted data for the source from circular regions of 15′′ and 30′′ centered on the pulsar position. The total counts encompassed in these regions are 391 and 1143 respectively, with the background con- tributing 151 and 562 counts, respectively. In order to search for an X-ray modulation at the PSR B0355+54 spin frequency, we first determined a predicted pulse frequency at the epoch of our XMM- Newton observations, assuming a linear spin-down rate and using the radio measurements (Hobbs et al. 2004; Manchester et al. 2005). We calculate f = 6.3945388 Hz at the midpoint of our observation (MJD 52315.7). As glitches and/or deviations from a linear spin-down may alter the period evolution, we then searched for a pulsed signal over a wider frequency range centered on f = 6.39454 Hz. We searched for pulsed emission using Fig. 5. -- The Chandra spectrum of the core emission of PSR B0355+54 with best-fit blackbody plus power-law model (top) and magnetized, pure H atmospheric plus power-law model (bot- tom). Also shown are the residuals from comparison of the data to the model in each case. 6 McGowan et al. Fig. 7. -- Z 2 1 -test (top) and Maximum Likelihood Periodogram (MLP; bottom) for the PN data of PSR B0355+54. The dom- inant peak from the Z 2 1 -test and the corresponding peak in the MLP are marked. The peaks occur at 6.3945447+0.0000167 −0.0000107 Hz and 6.3945467+0.00000821 −0.00000818 Hz, respectively. The dotted lines represent the 68% (χ2 = 2.71) and 90% (χ2 = 1.0) confidence levels for the frequencies in the MLP. The most significant Z 2 The frequency search of the data extracted from an aperture of radius 30′′ does not yield any significant peaks near to the predicted frequency with either search method. From our spatial analysis we know that the core of PSR B0355+54's emission lies within ∼ 5′′ of the pul- sar position. We have therefore also searched for pulsed modulations in data extracted from a smaller aperture, however we have used a radius of 15′′ as any smaller does not encompass enough counts for a meaningful analysis. n-statistic occurs for n = 1. With the number of harmonics equal to one, the Z 2 n- statistic corresponds to the well known Rayleigh statis- tic. We find three peaks with > 90% significance in the MLP, all with corresponding peaks from the Z 2 1 -test (see Figure 7). The dominant peak from the Z 2 1 -test occurs at 6.3945447+0.0000167 −0.0000107 Hz, with the corresponding peak in the MLP occurring at 6.3945467+0.00000821 −0.00000818 Hz. The uncertainties quoted are the 68% confidence limits on the position of the peak. Both frequencies are consis- tent, within the 68% contour, with the predicted pulse frequency, and with each other within the 90% contour. The second most prominent peak from the Z 2 1 -test, and the corresponding peak in the MLP, are not consistent with the predicted pulse frequency. While we have detected a frequency that is in agreement with the predicted pulse frequency for PSR B0355+54 we caution that the Z 2 1 peak has a prob- ability of chance occurrence of 3×10−3. Further observa- tions of the source are needed to show whether the mod- ulation detected is in fact pulsed X-ray emission from PSR B0355+54. We have folded the data on the pre- dicted pulse frequency and the frequency found from the Z 2-test (see Figure 8); by fitting the profiles with a si- nusoid we find that the modulation amplitude for the former is 25 ± 7% and 21 ± 8% for the latter. 6. DISCUSSION Our spatial analysis of the XMM-Newton and Chandra observations of PSR B0355+54 have not only revealed X- rays from the pulsar, but have provided definitive proof of Fig. 6. -- Spectral fitting to Chandra data of PSR B0355+54. First -- third panels: diffuse emission from regions 1 -- 3, respectively, with best fit power-law models. Also shown are the residuals from comparison of the data to the model in each case. two methods. In the first method we implement the Z 2 n test (Buccheri et al. 1983), with the number of harmon- ics n being varied from 1 to 5. In the second method we calculate the Rayleigh statistic (de Jager 1991; Mardia 1972) and then calculate the maximum likelihood peri- odogram (MLP; see e.g. Zane et al. 2002) using the C statistic (Cash 1979) to determine significant periodici- ties in the data sets. The Pulsar Wind Nebula of PSR B0355+54 7 RN S = 7.2+7.2 −2.2 km. Taking into account the possible detection of pulsed emission from PSR B0355+54 it is likely that the emitting region is a hot polar cap. It is suggested that the presence of a PWN is re- lated to the spin-down power of the pulsar, and for sources with log E . 36 the PWN emission efficiency is significantly reduced (Frail & Scharringhausen 1997; Gaensler et al. 2000; Gotthelf 2003). For PSR B0355+54 log E = 34.6, making it one of a handful of sources with spin-down power below this limit with a detectable PWN (cf. Geminga, Caraveo et al. 2003). Using the results from the blackbody plus power-law fit to the core emis- sion detected with Chandra we determine an isotropic un- absorbed luminosity in the 0.3 -- 10 keV band of 8.3 × 1030 erg s−1. With E = 4.5 ×1034 erg s−1 for PSR B0355+54, this leads to a conversion efficiency of 2 × 10−4. So in fact, we find that the conversion efficiency of the point source is similar to the values found for other pulsars (see e.g. Becker & Trumper 1997; Gaensler et al. 2004). Our analysis also indicates that the compact diffuse com- ponent is more luminous than the point source, with a conversion efficiency of 5 × 10−4 in the 0.3 -- 10 keV range. This result is again consistent with other sources (Becker & Trumper 1997). In addition, it is reported that when the pulsar spin-down energy is log E . 36.5 the morphology of the PWN seems to transition from toroidal to a jet/tail (Kaspi et al. 2005). Our measure- ments of PSR B0355+54 appear to agree with this trend. The morphology of the diffuse emission depends on how the interaction with the interstellar medium (ISM) or supernova remnant constrains the flow of particles (e.g. Reynolds & Chevalier 1984). For a pulsar that is moving with a supersonic space velocity, the interaction of the supersonic flow with the ambient medium causes the speed of the flow to decrease sharply, while the den- sity increases, forming a bow shock. In addition to the bow shock, which is at some distance ahead of the pul- sar, a reverse shock is formed nearer to the source which terminates the pulsar wind. The results of the spatial analysis of the XMM-Newton PN and Chandra ACIS data of PSR B0355+54 indi- cate that the bulk of the diffuse emission extends ∼ 50′′ [0.25(d/1.04 kpc) pc] downstream from the pulsar. Us- ing the measurements of the pulsar's proper motion (Chatterjee et al. 2004) we find that the transverse ve- locity of PSR B0355+54 is vt = 61 km s−1. This implies that the time taken for the pulsar to have traversed the length of the diffuse emission is > 4000 yr. In addi- tion, by considering the analysis of the XMM-Newton MOS1 data we find that the diffuse emission may ex- tend as far as ∼ 5′ [1.51(d/1.04 kpc) pc] from the point source. This results in a travel time of > 24000 yr for the pulsar. Following the work of Wang & Gotthelf (1998) (see also Kaspi et al. 2001) the synchrotron life- time of an electron of energy E (in keV) can be defined as ts ∼ 40E−1/2B−3/2 is the magnetic field in units of 10−4 G. Assuming that the dominant loss mechanism is synchrotron emission, i.e. B > 3.2 µG, and that the energy of the photon is E ∼ 5 keV, then ts ∼ 3000 yr. This indicates that the diffuse emis- sion that we detect is not due to particles deposited by the pulsar as it traveled through space. Hence, there yr, where B−3/2 −4 −4 Fig. 8. -- PN data in the 0.3 -- 10 keV energy range for PSR B0355+54 folded on the frequency predicted from the radio measurements (top) and the frequency found from the Z 2 1 -test (bot- tom). In both cases the data are folded using the radio ephemeris. diffuse emission extending in the opposite direction to the pulsar's proper motion. Similar detections of extended emission have been seen for other sources (e.g. N157B, Wang & Gotthelf 1998; PSR B1757-24, Frail & Kulkarni 1991; Kaspi et al. 2001; PSR B1957+20, Stappers et al. 2003; PSR B1951+32, Li et al. 2005), and have been interpreted as emission from a ram-pressure confined PWN. We cannot separate the core and diffuse emission com- ponents for the XMM-Newton data and find that the spectrum can be well-fitted with a power-law model with index Γ = 1.5+0.5 −0.3, similar to the value found for other PWNe (Kaspi et al. 2005). The nebular emission is most likely dominating the spectrum. The core emission from the Chandra data can be well-fitted by a thermal plus power-law model. A fit with a blackbody plus power-law model gives T = (2.32+1.16 −0.81) × 106 K and Γ = 1.0+0.2 −1.0. The fitted blackbody flux corresponds to an emitting ra- dius of 0.12+0.16 −0.07 km, where the distance to the source is 1.04 kpc. In this case, the size of the emitting re- gion implies that the flux originates from a hot polar cap. We can also fit the spectrum with a pure H, mag- netized atmospheric plus power-law model. With the distance to the source fixed at 1.04 kpc, this fit re- sults in T ∞ −0.4 and −0.22) × 106 K, Γ = 1.5+0.5 ef f = (0.45+0.20 8 McGowan et al. must be a constant supply of wind particles traveling at velocities greater than the space velocity of the pulsar. In addition, the particle flow velocity must be high enough such that the time for the flow to cross the length of the diffuse emission is less than the radiative lifetime of the particles. Using the Chandra data we have modeled the spectrum of the compact diffuse emission, excluding the contribu- tion from the pulsar, finding that the data can be well- fitted with a power-law. In other sources the power-law is seen to soften as one moves away from the pulsar posi- tion (see e.g. Slane et al. 2002; Li et al. 2005; Kaspi et al. 2005). An increase in the spectral index is expected as the particles will be cooler, i.e. older, at greater dis- tance from the pulsar. Our results indicate that we are detecting relatively hard emission, but due to the uncer- tainties, we are unable to comment on any changes in the spectral slope. To measure cooling the PWN must be of an adequate size, it may be that for PSR B0355+54 the compact diffuse region is not large enough for a substan- tial change in power-law index to be measured, and there are too few counts in the more extended diffuse region to perform a spectral analysis. It is noted however that by comparing the spectral indices from the blackbody plus power-law fit to the core emission and the power-law fit to the compact diffuse emission we do detect an increase in Γ of ∼ 0.5. Gaensler et al. (2004) have presented a detailed analy- sis of the diffuse X-ray emission associated with the radio source G359.23-0.82, also known as "the Mouse". Their hydrodynamic simulations show that there are a num- ber of regions that can be defined in a pulsar bow shock. These include a pulsar wind cavity, shocked pulsar wind material, contact discontinuity (CD) and shocked ISM. The energetic shocked particles from the pulsar are confined by the CD, the position of which denotes the transition to the shocked ISM. Following the method of Gaensler et al. (2004) we have estimated the distance be- tween the peak of the emission from PSR B0355+54 and the sharp cut-off in brightness ahead of the pulsar. Us- ing the same limit as Gaensler et al. (2004) i.e. where the X-ray surface brightness falls by 1/e2 = 0.14, we find a distance of 0.9′′ ± 0.2′′, giving the CD a projected radius of rCD = 0.004 ± 0.001 pc. Here, and in the fol- lowing, we have used a distance to the pulsar of 1.04 kpc (Chatterjee et al. 2004). From Eq. (1) of Gaensler et al. (2004) we can estimate the radius of the forward termi- nation shock (TS), rF TS ∼ 0.003 pc. This corresponds to an angular distance of θ = 0.59′′. Comparing our val- ues to those for the Mouse implies that the emission in front of the pulsar is more compact in PSR B0355+54 than for the Mouse. In both cases the close proximity of the forward TS to the peak X-ray emission renders the TS undetectable. Using our results and Eq. (2) of Gaensler et al. (2004) we find that PSR B0355+54 pro- duces a ram pressure of ρv2 t ∼ 1.4 × 10−9 ergs cm−3. Assuming cosmic abundances, this gives vt ∼ 247n−1/2 km s−1, where n0 is the number density of the ambient medium, so for PSR B0355+54 we determine n0 ≈ 0.06 cm−3, which is not unrealistic. 0 Additional information can be obtained by equating the pressure of the pulsar wind (assumed isotropic), to that of the ambient medium. By introducing the Mach number M = vt/cs, where cs is the adiabatic sound speed in the ambient medium, and using the same prescription as Gaensler et al. (2004) for a representative ISM pres- sure (i.e. PISM = 2400kP0 erg cm−3, with 0.5 . P0 . 5 and k is the Boltzmann's constant), this gives: E/[4π(rF T S)2c] = 2400kγISM P0M 2 , (1) from which we can obtain an estimate of the Mach num- ber. We find that for PSR B0355+54 the sound speed cs of the medium lies in the range 1 − 30 km s−1. The three principal phases of the ISM are generally named cold, warm and hot and are characterized by typical sound speed values of 1, 10 or 100 km s−1; according to this de- nomination our result implies that the pulsar is moving in either a cold or mildly warm ambient gas. For com- parison, in the case of the Mouse, Gaensler et al. (2004) found that the most probable pulsar velocity requires that the pulsar is moving through a warm phase of the ISM. TS ≫ rF Gaensler et al. (2004) also discuss the possible detec- tion of the backward TS in their data. Their simulations show that this feature has a closed surface, while the CD and bow shock are unrestricted. The backward TS should lie much further away from the pulsar than the forward TS, i.e. rB TS. In principle this means that the backward TS may be detectable. The possible dip we see in the profiles for the PSR B0355+54 data could indicate the presence of the backward TS. The angular separation of the dip in our data is ∼ 10′′, a value con- sistent with that for the Mouse. For a backward TS, Gaensler et al. (2004) predict that there would be a lack of spectral evolution, a result we have found for the dif- fuse emission of PSR B0355+54. However, we note that the feature in the Mouse data (and simulations) is quite compact in the north-south direction in comparison to the PSR B0355+54 feature. In addition, the number of counts we detect for PSR B0355+54 may hinder our investigation of the presence of such a feature. Deeper observations are needed to probe further the PWN of PSR B0355+54. A few days before submission of this paper, Tepedelen- lioglu & Ogelman submitted a paper based on the same (public) observations to ApJL (astro-ph/0512209). The authors wish to thank the referee for useful com- ments that have helped improve the paper. This work is based on observations obtained with XMM-Newton, an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA. Sup- port for this work was provided by the National Aeronau- tics and Space Administration through Chandra Award Number NNG04EF62I issued by the Chandra X-ray Ob- servatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of the Na- tional Aeronautics Space Administration under contract NAS8-03060. SZ thanks PPARC for its support through a PPARC Advanced Fellowship Becker, W., Trumper, J. 1997, A&A, 326, 682 REFERENCES Bhat, P. N., Acharya, B. S., Gandhi, V. N., Ramana Murthy, P. V., Sathyanarayana, G. P., Vishwanath, P. R. 1990, A&A, 236, 1 The Pulsar Wind Nebula of PSR B0355+54 9 Buccheri, R., et al. 1983, A&A, 128, 245 Caraveo, P. A., Bignami, G. F., DeLuca, A., Mereghetti, S., Pellizzoni, A., Mignani, R., Tur, A., Becker, W. 2003, Science, 301, 1345 Cash, W. 1979, ApJ, 228, 939 Chatterjee, S., Cordes, J. M., Vlemmings, W. H. T., Arzoumanian, Z., Goss, W. M., Lazio, T. J. W. 2004, ApJ, 604, 339 de Jager, O. C. 1991, ApJ, 378, 286 Frail, D. A., Kulkarni, S. R. 1991, Nature, 352, 785 Frail, D. A., Scharringhausen, B. R. 1997, ApJ, 480, 364 Gaensler, B. M., Stappers, B. W., Frail, D. A., Moffett, D. A., Johnston, S., Chatterjee, S. 2000, MNRAS, 318, 58 Gaensler, B. M. 2001, in Young Supernova Remnants, eds. S. S. Holt & U. Hwang, AIP Conference Proceedings, 565, 295 Gaensler, B. M., van der Swaluw, E., Camilo, F., Kaspi, V. M., Baganoff, F. K., Yusef-Zadeh, F., Manchester, R. N. 2004, ApJ, 616, 383 Gotthelf, E. V. 2003, ApJ, 591, 361 Helfand, D. J. 1983, in IAU Symp. 101, Supernova Remnants and Their X-ray Emission, eds. J. Danzinger & P. Gorenstein, Dordrecht: Reidel, 471 Hobbs, G., Lyne, A. G., Kramer, M., Martin, C. E., Jordan, C. 2004, MNRAS, 353, 1311 Kaspi, V. M., Gotthelf, E. V., Gaensler, B. M., Lyutikov, M. 2001, ApJ, 562, 163 Kaspi, V. M., Roberts, M. S. E., Harding, A. K. 2005, in Compact Stellar X-ray Sources, ed. W. H. G. Lewin & M. van der Klis (Cambridge: Cambridge Univ. Press), in press Li, X. H., Lu, F. J., Li, T. P. 2005, ApJ, 628, 931 Manchester, R. N., Hobbs, G. B., Teoh, A., Hobbs, M. 2005, AJ, 129, 1993 Mardia, K. V. 1972, Statistics of Directional Data (London: Academic) Pavlov, G. G., Shibanov, Y. A., Zavlin, V. E., Meyer, R. D. 1995, in The Lives of Neutron Stars eds. A. Alpar, U. Kiliz´oglu & J. van Paradijs, Kluwer Academic Publishers, p. 71 Pavlov, G. G., Kargaltsev, O. Y., Sanwal, D., Garmire, G. P. 2001, ApJ, 554, 189 Rees, M. J., Gunn, J. E. 1974, MNRAS, 167, 1 Reynolds, S. P., Chevalier, R. A., 1984, ApJ, 278, 630 Seward, F. D., Wang, Z.-R. 1988, ApJ, 332, 199 Slane, P. 1994, ApJ, 437, 458 Slane, P. O., Helfand, D. J., Murray, S. S. 2002, ApJ, 571, 45 Stappers, B. W., Gaensler, B. M., Kaspi, V. M., van der Klis, M., Lewin, W. H. G. 2003, Sci, 299, 1372 Struder, L., et al. 2001, A&A, 365, L18 Tepedelenlioglu, E., (astro-ph/0512209) Ogelman, H. 2005, ApJL, Turner, M. J. L., et al. 2001, A&A, 365, 27 Wang, Q. D., Gotthelf, E. V. 1998, ApJ, 494, 623 Weisskopf, M. C. 2000, ApJ, 536, 81 Zane, S., et al. 2002, MNRAS, 334, 345 submitted
astro-ph/0511209
1
0511
2005-11-08T05:40:11
Analysis of MERCATOR data Part I: variable B stars
[ "astro-ph" ]
We re-classified 31 variable B stars which were observed more than 50 times in the Geneva photometric system with the P7 photometer attached to the MERCATOR telescope (La Palma) during its first 3 years of scientific observations. HD89688 is a possible beta Cephei/slowly pulsating B star hybrid and the main mode of the COROT target HD180642 shows non-linear effects. The Maia candidates are re-classified as either ellipsoidal variables or spotted stars. Although the mode identification is still ongoing, all the well-identified modes so far have a degree l = 0, 1 or 2.
astro-ph
astro-ph
Comm. in Asteroseismology Vol. 144, 2003 Analysis of mercator data Part I: variable B stars P. De Cat1,2, M. Briquet2, 6, C. Aerts2,3, K. Goossens2, S. Saesen2, J. Cuypers1, K. Yakut2, R. Scuflaire4, M.-A. Dupret5 and many observers 1 Royal Observatory of Belgium, B-1180 Brussel, Belgium 2 Instituut voor Sterrenkunde, Katholieke Universiteit Leuven, B-3001 Leuven, Belgium 3 Department of Astrophysics, Radboud University Nijmegen, 6500 GL Nijmegen, the Netherlands 4 Institut d'Astrophysique et de G´eophysique, Universit´e de Li`ege, B-4000 Li`ege, Belgium 5 Observatoire de Paris, LESIA, 92195 Meudon, France Abstract We re-classified 31 variable B stars which were observed more than 50 times in the Geneva photometric system with the p7 photometer attached to the mercator telescope (La Palma) during its first 3 years of scientific observations. HD 89688 is a possible β Cephei/slowly pulsating B star hybrid and the main mode of the corot target HD 180642 shows non-linear effects. The Maia candidates are re-classified as either ellipsoidal variables or spotted stars. Although the mode identification is still ongoing, all the well-identified modes so far have ℓ ≤ 2. Introduction The mercator telescope is a 1.2-m telescope located on the Roque de los Muchachos observatory on La Palma (Spain). Since the start of scientific observations in spring 2001, this telescope has been intensively used to observe variable B, A and F main sequence stars with the p7 photometer, providing quasi-simultaneous observations in the 7 passbands of the Geneva photometric system. The first results obtained after 18 months of observations were already presented by De Cat et al. (2004). We now present results after 3 years of collecting data. In Part I (this paper), the analysis of the 9023 datapoints of variable B stars is discussed while Paper II (Cuypers et al., these proceedings) focuses on the analysis of the 5149 datapoints of variable A and F stars. (2004) and De Ridder et al. We here restrict ourselves to the 31 variable B stars which were not included in multi-site campaigns and which were observed at more than 50 epochs. Based on the photometric observations gathered with the satellite mission hipparcos, these objects were previously classified as candidate β Cephei stars (β Ceps; main-sequence B 0 -- 3 stars pulsating in low order, low degree p/g-modes with periods of 3 -- 8 h), slowly pulsating B stars (SPBs; main- sequence B 3 -- B 9 stars pulsating in high order, low degree g-modes with periods of 0.5 -- 3 d) and Maia stars (Maias; variable main-sequence stars situated between the SPBs and the δ Scuti stars). They are respectively given with squares, triangles and stars in Fig. 1. 6Postdoctoral Fellow of the Fund for Scientific Research, Flanders 2 Analysis of mercator data Part I: variable B stars Figure 1: Position in the H-R diagram of the 31 variable B stars discussed in this paper. The candidate β Cephei stars, slowly pulsating B stars and Maia stars are respectively given with squares, triangles and stars. HD 89688 is given with a full symbol. The ZAMS and TAMS are given with dashed lines, and the theoretical instability strips for β Cephei and slowly pulsating B stars, as given by Pamyatnykh (1999), with full lines. Frequency analysis The time series in the Geneva passbands and colours were both subjected to a detailed fre- quency analysis with the PDM (Stellingwerf 1978) and Lomb-Scargle (Scargle 1982) methods. Since our ground-based data-sets suffer from strong aliasing, the space-based photometric observations of the hipparcos satellite proved to be very useful to extract the physical fre- quencies in some cases. Our results enable us to re-classify the stars into the following categories by using the same criteria as De Cat et al. (2004): • SPBs: 11 multiperiodic (HD 1976, 3379, 21071, 25558, 28114, 28475, 179588, 182255, 191295, 206540, 222555), and 2 monoperidic (HD 138003, 208057) • β Ceps: 6 multiperiodic (HD 13745, 13831, 14053, 21803, 180642, 203664) • Hybrid star: HD 89688 • Spotted stars: HD 46005, 154689, 169820 • Ellipsoidal variables: HD 24094, 112396, 149881, 208727 • Be star: HD 180968 • Constant stars: HD 19374, 214680, 217782 For all the periodograms and phase diagrams, we refer to De Cat et al. (in preparation). For HD 89688, we now have marginal evidence for the SPB-like frequency 0.7965(6) d−1 (or one of its aliases), while the hipparcos photometry points towards β Cep-like frequency 7.3902(5) d−1. Its position in the H-R diagram is compatible with the classification as a hybrid star (full symbol in Fig. 1). For the multiperiodic corot target HD 180642, the first mode is a high amplitude mode which shows non-linear effects. We detect up to the second harmonic of ν1 = 5.486971(6) d−1, making it only the second β Cep star for which more than one harmonic is observed (Aerts et al., in preparation). Note that for all the Maias, i.e. HD 46005, 154689, 169820, 208727, the observed variations can be explained by mechanisms other than pulsations. Mode identification For the mode identification, we applied the method of the photometric amplitudes (Dupret et al. 2003) in which the observed and theoretical amplitude ratios relative to the amplitude P. De Cat, M. Briquet, C. Aerts et al. 3 Figure 2: Presentation of the grids of equilibrium models used for the mode identification of the SPBs. The evolution tracks of grid 1 and 2 (see text) with masses between 2 -- 8 M⊙ in steps of 0.1 M⊙ are respectively given in light and dark grey. The boxes represent the observed errorboxes of the 13 SPBs in our sample. The filled box corresponds to HD 179588. (2005). in the U filter are compared. For the SPBs, we confronted the results based on 2 grids of equilibrium models (Fig. 2). Grid 1 consists of models calculated with CLES-013 (written by R. Scuflaire) with an initial mass fraction of metals Z0 = 0.020 and of hydrogen X0 = 0.70, mixing-length αconv = 1.80, and the standard metal mixture of Grevesse & Noels (1993). Grid 2 was obtained with CLES-018.2 with the 'new' solar values Z0 = 0.015, X0 = 0.71, αconv = 1.75 and the standard metal mixture of Asplund et al. In both cases, we used the CEFF equation of state (Christensen-Dalsgaard & Dappen 1992) and a Kurucz atmosphere with the junction point at optical depth τ = 10, and we assumed neither convective overshooting nor diffusion. One of the main changes between CLES-013 and CLES-018.2 is the use of the new value of the cross section of 14N (p, γ)15O recently measured by Formicola et al. (2004). We calculated the non-adiabatic eigenfunctions and eigenfrequencies for g-modes with ℓ ≤ 3 with the code MAD (written by M.-A. Dupret). For each star, we selected the models within the observed errorbox of log(Tef f ) and log g (boxes in Fig. 2), and selected the eigenfrequency which is the closest to the observed frequency to calculate the theoretical amplitude ratios. In Fig. 3, we give a representative example of our results, i.e. for the two main modes of HD 179588. Although there are significant differences in the position and/or the shape of the theoretical curves for the higher degree modes of grid 1 (left) and 2 (right), the identification of the modes remains the same, i.e. ℓ = 1 or 2 for the mode corresponding to ν1 = 0.856543(15) d−1 (top), and ℓ = 1 for the mode corresponding to ν2 = 2.04263(5) d−1 (bottom). In general, these differences coming from the use of 2 different grids increase for increasing values of the observed frequency. So far, all the well-identified SPB modes have ℓ = 1 or 2. For the β Ceps, the mode identification is still ongoing. Conclusions Our photometric survey allowed a significant contribution in the classification of variable B stars. HD 89688 is a possible β Cep/SPB hybrid star and the corot target HD 180642 is a multiperiodic β Cep star of which the main mode shows non-linear effects (Aerts et al., in preparation). Amongst the 31 targets with a sufficient amount of data, we identified 4 ellip- soidal variables and 4 spotted stars. Their classification should be checked by supplementary spectroscopic observations. The mode identification is still ongoing, but all well-identified 4 Analysis of mercator data Part I: variable B stars Figure 3: Photometric mode identification for ν1 = 0.856543(15) d−1 (top) and ν2 = 2.04263(5) d−1 (bottom) of HD 179588. For each theoretical model within the observed range of log(Tef f ) and log g, the theoretical amplitude ratios for modes with ℓ = 1, 2 and 3 are represented with an individual line. The dots indicate the observed amplitude ratios and their standard error. The left and right panels show the results obtained with grid 1 and 2 respectively (see text). modes have ℓ ≤ 2 so far. The final results of our survey will be given by De Cat et al. (in preparation). The merca- tor observations allow to take the first steps in asteroseismic modeling for two multiperiodic β Cep stars, i.e. HD 203664 (Aerts et al., submitted to A&A) and HD 21803 (Saesen et al., in preparation). For 12 Lac and V2052 Oph, the mercator telescope was included in multi-site campaigns. The data of these objects are being analysed by Handler et al. (submitted to MNRAS) and Handler et al. (in preparation) respectively. Acknowledgments. This work is based on observations collected with the p7 photometer attached to the mercator telescope (La Palma, Spain). We are very much indebted to all the observers from Leuven University. CA and JC acknowledge support from the Fund for Scientific Research (FWO) - Flanders (Belgium) through project G.0178.02. References Asplund M., Grevesse N., Sauval A.J., 2005, ASP Conf. Ser. 336, 25 Christensen-Dalsgaard J., Dappen W., 1992, A&ARv 4, 267 De Cat, P., De Ridder, J., Uytterhoeven, K., et al., 2004, ASP Conf. Proc. 310, 238 De Ridder, J., Cuypers, J., De Cat, P., et al., 2004, ASP Conf. Proc. 310, 263 Dupret M.-A., De Ridder J., De Cat P., et al., 2003, A&A 398, 677 Formicola A., Imbriani G., Costantini H., et al., 2004, Physics Letters B 591, 61 Grevesse N., Noels A., 1993, Physica Scripta 47, 133 Pamyatnykh, A.A., 1999, Acta Astr. 49, 119 Scargle, J.D. 1982, ApJ 263, 835 Stellingwerf, R.F. 1978, ApJ 224, 953
astro-ph/0611461
1
0611
2006-11-15T01:28:49
Understanding white dwarf binary evolution with white dwarf/main sequence binaries: first results from SEGUE
[ "astro-ph" ]
Close white dwarf binaries make up a wide variety of objects such as double white dwarf binaries, which are possible SN Ia progenitors, cataclysmic variables, super soft sources, or AM CVn stars. The evolution and formation of close white dwarf binaries crucially depends on the rate at which angular momentum is extracted from the binary orbit. The two most important sources of angular momentum loss are the common envelope phase and magnetic braking. Both processes are so far poorly understood. Observational population studies of white dwarf/main sequence binaries provide the potential to significantly progress with this situation and to clearly constrain magnetic braking and the CE-phase. However, the current population of white dwarf/main sequence binaries is highly incomplete and heavily biased towards young systems containing hot white dwarfs. The SDSSII/SEGUE collaboration awarded us with 5 fibers per plate pair in order to fill this gap and to identify the required unbiased sample of old white dwarf/main sequence binaries. The success rate of our selection criteria exceeds 65% and during the first 10 months we have identified 41 new systems, most of them belonging to the missed old population.
astro-ph
astro-ph
**FULL TITLE** ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION** **NAMES OF EDITORS** Understanding white dwarf binary evolution with white dwarf/main sequence binaries: first results from SEGUE M.R. Schreiber1 , 2, A. Nebot Gomez-Moran2, A.D. Schwope2 1 Universidad de Valparaiso, Facultad de Ciencias, Departamento de Fisica y Meteorologia, Av. Gran Bretana 1111, Valparaiso, Chile 2 Astrophysikalisches Inst. Potsdam, An der Sternwarte 16, 14482 Potsdam, Germany Abstract. Close white dwarf binaries make up a wide variety of objects such as double white dwarf binaries, which are possible SN Ia progenitors, cataclysmic variables, super soft sources, or AM CVn stars. The evolution and formation of close white dwarf binaries crucially depends on the rate at which angular mo- mentum is extracted from the binary orbit. The two most important sources of angular momentum loss are the common envelope phase and magnetic braking. Both processes are so far poorly understood. Observational population stud- ies of white dwarf/main sequence binaries provide the potential to significantly progress with this situation and to clearly constrain magnetic braking and the CE-phase. However, the current population of white dwarf/main sequence bina- ries is highly incomplete and heavily biased towards young systems containing hot white dwarfs. The SDSSII/SEGUE collaboration awarded us with 5 fibers per plate pair in order to fill this gap and to identify the required unbiased sam- ple of old white dwarf/main sequence binaries. The success rate of our selection criteria exceeds 65% and during the first 10 months we have identified 41 new systems, most of them belonging to the missed old population. 1. Introduction Close binaries containing at least one white dwarf span a wide range of interest- ing and exotic stars, such as detached white dwarf binaries, cataclysmic variables (CVs), or AM CVn binaries. Besides offering the opportunity to study physics under extreme conditions, these objects are extremely important in the gen- eral astrophysical context: Supernova Ia arise either from merging binary white dwarfs or from interacting white dwarf/main sequence binaries and AM CVn stars are expected to significantly contribute to the gravitational wave back- ground which will be measured by LISA. All the different types of close white dwarf binaries have two points in common: (1) they evolved through at least one common envelope (CE) phase and (2) they undergo subsequent orbital angular momentum loss (AML). Sad but true, the physics of both the CE and AML are very poorly understood. In current theories the CE phase is simply approximated by a parameter- ized energy (Paczy´nski 1976; Webbink 1984; Willems & Kolb 2004) or angular momentum equation (Nelemans & Tout 2005). Both descriptions differ signifi- cantly in the predicted outcome of the CE phase and in both prescriptions the efficiency to "use" the orbital energy (angular momentum) to expel the enve- 1 2 lope is very uncertain. Hence, the CE phase is probably the least understood period of close binary evolution. Once the envelope is expelled, the evolution of the post common envelope binary (PCEB) is mainly driven by AML due to magnetic braking. Unfortunately, the two currently favoured prescriptions for magnetic braking (Verbunt & Zwaan 1981; Andronov et al. 2003), differ by up to two orders of magnitude. Even worse, it is not clear whether magnetic brak- ing is continuously present or if it gets disrupted when the secondary star is fully convective. In order to explain the orbital period gap observed in the period distribution of CVs, one needs to assume the latter (e.g. King 1988; Howell et al. 2001) while observations of single low mass stars do not show any evidence for such a discontinuity (e.g. Pinsonneault et al. 2002). Significant progress in the theoretical modelling of the CE phase and AML due to magnetic braking will clearly need observational input. A quantitative test of the current theories requires the knowledge of a large and unbiased pop- ulation of close binaries that underwent a CE and subsequent orbital AML. The ideal class of stars to provide such observational constraints on the CE and magnetic braking models are detached PCEBs consisting of a white dwarf and a main sequence star, as (1) white dwarf binaries are intrinsically numerous, (2) the properties of both stellar components are well-understood, and (3) they have rather short orbital periods (∼ 2h−50 d). 2. PCEBs in the pre-SDSS era Schreiber & Gansicke (2003) analysed the population of PCEBs with determined orbital period and white dwarf temperature. Their sample consisted of only 30 systems -- a surprisingly small number when compared with the more than 1000 CVs listed in Downes et al. (2006). Even worse, the detailed analysis of Schreiber & Gansicke (2003) showed that the small sample of 30 PCEBs is also heavily biased towards hot white dwarfs and late type secondary star spectral types. This bias is a natural consequence of the way PCEBs have been discovered in the past: as white dwarfs in the first place, with some evidence for a faint red companion found later. Finally, Schreiber & Gansicke (2003) calculated the evolutionary time scale of the 30 young (containing hot white dwarfs) PCEBs and find that most of them have passed only a very small fraction of their PCEB lifetime. This immediately leads to the prediction of a large population of old PCEBs containing cold white dwarfs which has not yet been identified. 3. The biases of the SDSS DR4 sample Since the first data release of the Sloan Digital Sky Survey (SDSS), the sit- uation changed drastically. Based on SDSS imaging and some DR 1 spectra Smolci´c et al. (2004) identified a new stellar locus, i.e. the white dwarf/main sequence (WD/MS) binary bridge. The population of these WD/MS binaries consists of wide binaries that will never interact and whose components evolve like single stars and close binaries that went through a common envelope phase (PCEBs). The SDSS turned out to be also very efficient in spectroscopically identifying new unresolved WD/MS binaries. Recently Silvestri et al. (2006) published a list 3 of ∼ 747 new WD/MS binary systems found in SDSS/DR4. However, as stated by Silvestri et al. (2006) themselves, the SDSS DR4 sample is again subject to strong observational biases. The WD/MS systems identified in SDSS/DR4 originate primarily from two different channels: the colour selection described in Silvestri et al. (2006) and serendipitous objects from QSO fibres. The color selection used by Silvestri et al. (2006) selects hot systems mainly because of the cut used in u − g versus g − r and the SDSS QSO selection algorithm (see Richards et al. 2002, Fig. 7) explicitly excludes the color-color space of cold white dwarf/main sequence binaries. Hence both channels produce predominantly WD/MS binaries with hot white dwarfs, i.e. young objects, which -- according to Schreiber & Gansicke (2003) -- represent only the minority of all WD/MS binaries. 4. Identifying old PCEBs with SEGUE A true constraint on AML mechanisms in close binaries will only be possible once a representative sample of PCEBs has been identified. As partners of SDSS II we are running a successful program (PI: M. Schreiber) identifying the missing cold WD/MS binary population. The SEGUE-collaboration awarded us with 5 fibers per SEGUE plate pair (∼ 7deg2) and we developed special color-cuts to select WD/MS systems containing cold white dwarfs, i.e. u − g < 2.25 g − r > −0.2 g − r < 1.2 r − i > 0.5 r − i < 2.0 g − r > −19.78 ∗ (r − i) + 11.13 g − r < 0.95 ∗ (r − i) + 0.5 i − z > 0.5 for r − i > 1.0 i − z > 0.68 ∗ (r − i) − 0.18 for r − i <= 1.0 15 < g < 20. The main selection criteria are shown in Fig. 1 as black lines. In the first 10 months 41 SEGUE-plates with WD/MS target selection have been observed. During the first drilling run in Oct. 2005, the above criteria have been applied to reddening corrected magnitudes. This led to the identification of several nearby single M-dwarfs whose reddening corrected colors resemble WD/MS binaries containing cold white dwarfs. The success rate for the Oct. 2005 plates therefore is only 14/35 = 40% on 22 plates. Since 2006 we use non-corrected ugriz magnitudes and our success rate increased to 27/40 = 67.5% on 19 plates. Fig. 1 shows the positions of the 75 SEGUE-WD/MS candidates including the 41 WD/MS systems (black open squares). Also shown are the Silvestri et al. (2006) sample (black points) and the QSO and single star population (gray). As an example for the 41 identified systems, Fig. 2 shows the SEGUE spectrum of one cool WD/MS binary. We determined the white dwarf temperature and the spectral type of the secondary of the 41 WD/MS binaries by fitting simultaneously the composite binary spectrum. The resulting distributions are shown in Fig. 3. Compared to the SDSS DR4 sample published by Silvestri et al. (2006) our sample contains significantly more WD/MS systems with cold white dwarfs and/or early type secondary stars thereby overcoming previous biases in the sample of known WD/MS binaries. 4 Figure 1. The SEGUE-WD/MS color cuts (black lines) in two color-color diagrams. Quasars and single stars are shown in grey. The (not reddening corrected) positions of the 75 WD/MS candidates selected during the first 10 months are marked as open squares. In the first drilling run we used reddening corrected magnitudes and some nearby M-dwarfs (those below the lower vertical lines) appeared as WD/MS candidates. Since 2006 we select our candidates without reddening correction and the success rate increased to 67.5%. The SDSS/DR4 WD/MS population (Silvestri et al. 2006) is shown as black points. Apparently, the overlap of the with the SEGUE selection is rather small as the latter is especially designed to identify the missing old population. Figure 2. color-color space (using fiber-magnitudes). The spectrum of a SEGUE WD/MS binary and the position in 5 The distributions of the effective temperature of the white dwarf Figure 3. (top) and spectral type of the main sequence secondary star (bottom). As predicted, compared to Silvestri et al. (2006, their Fig. 5 and 10) the SEGUE- WD/MS sample contains significantly more systems with cold white dwarfs and early type secondaries. According to the SEGUE baseline ∼ 200 plate pairs will be observed until mid 2008 and we therefore expect to identify ∼ 400 − 500 new WD/MS binaries by then. Together with the systems identified by Silvestri et al. (2006) the re- sulting more than ∼ 1200 WD/MS systems will form the data base to constrain theories of close binary evolution. 5. Constraining close binary evolution with PCEBs In principal the three big questions of close binary evolution can be answered using a large sample of PCEBs with known orbital period, secondary spectral type, and white dwarf temperature: (1) The disrupted magnetic braking sce- nario predicts an increase of the relative number of PCEBs by a factor ∼ 1.7 in the range of secondary spectral types M3-M5 (see Politano & Weiler 2006). To confirm or disprove the predicted increase one needs to identify PCEBs with M3-M5 secondaries among the WD/MS population. As the mean PCEB life- time can be rather large, the expected increase of the relative number of PCEBs will be more pronounced in the old SEGUE population. (2) The strength of AML can be estimated by comparing the orbital period distributions of PCEBs at different times of the PCEB evolution. A representative sample of PCEBs for secondary spectral types M0-M8 and effective temperatures of the white dwarf of Twd ∼ 10000 − 40000 K is required. (3) The predictions of the two currently favoured prescriptions of the CE phase differ in particular in the predicted or- bital period distribution of long orbital period PCEBs (Nelemans & Tout 2005). 6 Consequently, identifying the long orbital period end of the PCEB population will clearly constrain current theories of the CE phase. To sum up, characterizing a large sample of PCEBs provides the potential to solve the three most important problems in close binary evolution. To that end we have initiated a large-scale follow-up programme to identify and char- acterize the PCEBs among the WD/MS sample involving telescopes at both hemispheres and utilizing multi-epoch spectroscopy, time-resolved photometry, and astrometry with promising first results. Acknowledgments. We acknowledge support by the Deutsches Zentrum fur Luft- und Raumfahrt (DLR) under contract FKZ 50 OR 0404 (MRS, ANGM) and FONDECYT, grant 1061199 (MRS). Funding for the SDSS and SDSS-II has been provided by the A. P. Sloan Foundation, the Participating Institutions, the NSF, the U.S. Dept. of Energy, the Nat. Aeronautics and Space Admin., the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participating Institutions are the American Mu- seum of Nat. History, Astrophys. Inst. Potsdam, University of Basel, University of Cambridge, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scien- tist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck- Institute for Astrophysics (MPA), New Mexico State University, Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton Uni- versity, the United States Naval Observatory, and the University of Washington. References Andronov, N., Pinsonneault, M., & Sills, A. 2003, ApJ, 582, 358 Downes, R. A., Webbink, R. F., Shara, M. M., Ritter, H., Kolb, U., & Duerbeck, H. W. 2006, VizieR Online Data Catalog, 5123, 0 Howell, S. B., Nelson, L. A., & Rappaport, S. 2001, ApJ, 550, 897 King, A. R. 1988, QJRAS, 29, 1 Nelemans, G. & Tout, C. A. 2005, MNRAS, 356, 753 Paczy´nski, B. 1976, IAU Symp. 73: Struct. and Evolution of Close Binaries, 75 Pinsonneault, M. H., Andronov, N., & Sills, A. 2002, in The Physics of Cataclysmic Variables and Related Objects, ed. B. T. Gansicke, K. Beuermann, & K. Reinsch (ASP Conf. Ser. 261), 208 -- 216 Politano, M. & Weiler, K. P. 2006, ApJ, 641, L137 Richards, G. T., Fan, X., Newberg, H. J., & 18 co-authors, 2002, AJ, 123, 2945 Schreiber, M. R. & Gansicke, B. T. 2003, A&A, 406, 304 Silvestri, N. M., Hawley, S. L., West, A. A., & 20 co-authors, 2006, AJ, 131, 1674 Smolci´c, V., Ivezi´c, Z., Knapp, G. R., & 14 co-authors, 2004, ApJ Lett., 615, L141 Verbunt, F. & Zwaan, C. 1981, A&A, 100, L7 Webbink, R. F. 1984, ApJ, 277, 355 Willems, B. & Kolb, U. 2004, A&A, 419, 1057
astro-ph/9411017
1
9411
1994-11-03T20:14:00
The Cosmological Mass Distribution from Cayley Trees with Disorder
[ "astro-ph" ]
We present a new approach to the statistics of the cosmic density field and to the mass distribution of high-contrast structures, based on the formalism of Cayley trees. Our approach includes in one random process both fluctuations and interactions of the density perturbations. We connect tree-related quantities, like the partition function or its generating function, to the mass distribution. The Press \& Schechter mass function and the Smoluchowski kinetic equation are naturally recovered as two limiting cases corresponding to independent Gaussian fluctuations, and to aggregation of high-contrast condensations, respectively. Numerical realizations of the complete random process on the tree yield an excess of large-mass objects relative to the Press \& Schechter function. When interactions are fully effective, a power-law distribution with logarithmic slope -2 is generated.
astro-ph
astro-ph
THE COSMOLOGICAL MASS DISTRIBUTION FROM CAYLEY TREES WITH DISORDER A. CAVALIERE Astrofisica, Dipartimento di Fisica II Universit`a di Roma via della Ricerca Scientifica 1, I-00133 Roma (Italy) and N. MENCI SISSA, Via Beirut 2-4, I-34014 Trieste (Italy) permanent address: Osservatorio Astronomico di Roma Via dell'Osservatorio, I-00040 Monteporzio, Roma (Italy) 4 9 9 1 v o N 3 1 v 7 1 0 1 1 4 9 / h p - o r t s a : v i X r a preprint ROM2F/94/25, Ap. J. in press 1 ABSTRACT. We present a new approach to the statistics of the cosmic density field and to the mass distribution of high-contrast structures, based on the for- malism of Cayley trees. Our approach includes in one random process both fluc- tuations and interactions of the density perturbations. We connect tree-related quantities, like the partition function or its generating function, to the mass dis- tribution. The Press & Schechter mass function and the Smoluchowski kinetic equation are naturally recovered as two limiting cases corresponding to indepen- dent Gaussian fluctuations, and to aggregation of high-contrast condensations, respectively. Numerical realizations of the complete random process on the tree yield an excess of large-mass objects relative to the Press & Schechter function. When interactions are fully effective, a power-law distribution with logarithmic slope -2 is generated. Subject headings: cosmology: theory – galaxies: clustering – large scale structure of the universe – galaxies: formation 1. INTRODUCTION The mass distribution N (M ) of high-contrast structures constitutes a central link between observed extragalactic sources and the physics of the early universe. In the canonical scenario such structures form by direct hierarchical collapses (DHCs). In the expanding Universe small overdensities above the decreasing back- ground density ρ are weakly gravitationally unstable (see Peebles 1993). The con- trasts δ ≡ ∆ρ/ρ grow slowly in their linear regime, with δ ∝ t2/3 in a critical 2 universe. As the contrasts approach unity the perturbations detach from the Hub- ble expansion, turn around, and in a comparable time non linearly collapse to end up in high-contrast virialized structures. As for the "initial" conditions, the physics of the early universe (see Kolb & Turner 1990) suggests the perturbations started in the form of a Gaussian random field. Their power spectrum at z < 103 may be piecewise approximated by δk2 ∝ kn with −3 < n < 1. So nonlinear conditions k3δk2 ∼ 1 are reached sequentially at larger and larger sizes, with the rich clusters (Abell 1958) forming now. For the range of masses M ∼> 1012M⊙ the collapses are little affected by dissipation, and the DHC scenario gives rise yo a theory of elaborate elegance (see Peebles 1965; Gunn & Gott 1972; Press & Schechter 1974, hereafter PS; Rees & Ostriker 1977; Bond et al. 1991). The linear density field δ is smoothed or averaged at each point with a filter function of effective size R corresponding to a mass M ∝ ρR3. On each scale M , the variance of δ(M ) yields the dispersion σ ∝ M −a (a ≡ 1/2 + n/6) of the "initial" Gaussian distribution p(δ, σ). The nonlinear collapses are modeled after the pattern provided by the top-hat smoothing filter; namely, isolated spherical and homogeneous overdensities which virialize, with a definite characteristic mass Mc(t) ∝ t2/3a in a critical universe, by the time when the actual contrasts reach values ∼ 2 102. As this time would correspond to a nominal contrast δc ≈ 1.7 in the extrapolated linear behavior, such threshold is taken to separate the linear regime from bona fide condensations. 3 The condensed mass fraction is provided by the fraction of spheres where the linear δ(M ) crosses the threshold δc. The mass assigned to each collapsed object is twice that of the largest sphere wherein such a condition applies, while the smaller ones are disregarded. This elaborate selection rule can be written simply as N (M ) M dM = − 2ρ dR ∞ of m ≡ M/Mc(t) and σ ∝ m−a yields δc dδp(δ, σ), as originally proposed by PS, and in terms N (M, t) = ρ 2aδc √2πM 2 c (t)σ m−2e−δ2 c /2 σ2 . (1.1) The single parameter δc should comprise the complex nonlinear dynamics of the collapses. Once its value is set on the basis of, e.g., the top-hat model, the functional form of eq. 1.1 at any given t self-similarly depends on the initial spectrum only; specifically, on the spectral index n and on the amplitude taken by σ at the scale 8h−1 Mpc singled out by unit variance in galaxy counts. Self- similarity is stressed in the analysis of Bond et al. 1991, who examine the linear field at one epoch (say, the present to) on the resolution S ∝ M −2a ∝ σ2, and compare the resulting δ(S) with a lowering threshold 1.7 (t/to)−2/3. The same authors clarify in terms of excursion sets the statistics underlying eq. 1.1. They show that δ(S), when extracted from the Gaussian field using a sharp k-space filter, executes a simple (Markovian) random walk governed by a diffusion equation. The diffusive flux density per unit resolution of such trajectories, having their first up-crossing through the threshold δc within dM at M , yields the PS expression complete with shape and amplitude, in agreement with the selection rule. 4 For all its background, the result 1.1 has a number of drawbacks. The above prescriptions avoid overcounting substructures and imply a uniform timescale 3t/2 for all collapses. By the same token, however, they are likely to underplay sub- structures and to overestimate the normalization. In fact, computations which assign definite sizes and timescales to the collapses (Cavaliere, Colafrancesco & Scaramella 1991) yield longer permanence of substructures and a lower normaliza- tion for N (M, t). Equivalently, filters localized in the ordinary rather than in the k-space (Bond et al. 1991) change the mass assignment to the collapsing density peaks and yield different statistics and generally straighter shapes. Observations, especially in X-rays, image abundant substructures within clusters, ranging up to truly binary configurations (see Jones & Forman 1992; White, Briel, & Henry 1993). In addition, a considerable body of evidence indicates that the direct collapse scenario is incomplete at the high-mass end. For example, at galactic scales the cD's outnumber the Schechter luminosity distribution (Schechter 1976; Bhavsar 1989); the building up of their bodies is best understood in terms of aggregations of normal members in groups as described by the Smoluchowski kinetic equation (Cavaliere, Colafrancesco, & Menci 1992, hereafter CCM). Galaxy interactions and merging involving one bright partner are also likely to stimulate the emissions from active galactic nuclei at z ∼< 2.5, as indicated by statistics, by morphologies of the host galaxies, and by the richness of their environments (see Heckman 1993; Bahcall & Chokshi 1991); numerical experiments and theoretical studies 5 also concur to this view (Shlosman 1990; Barnes & Hernquist 1991). At larger scales, the imaging X-ray observations referred to above provide many snapshots of groups and clusters at various stages of essentially binary aggregations. In fact, high-resolution N-body experiments (e.g., Brainerd & Villumsen 1992; Katz, Quinn, & Gelb 1993; Jain & Bertschinger 1994) show structure formation to be a far more complex process than envisaged by the simple DHC scenario. It includes, in addition to direct collapses, frequent encounters and aggregations within clusters and within larger scale, precursor structures with the dimension- ality of sheets and filaments. The resulting N (M, t) shows, relative to eq. 1.1, a slower evolution and a different shape due to an excess at large M . Here we propose a novel approach to a satisfactory N (M, t). We discuss in §2 and §3 how the mass distribution is generated by a complete statistics in the resolution-contrast plane. This comprises as limiting cases both the direct collapses from independent Gaussian fluctuations (§4), and pure dynamical aggregations of high-contrast condensations (§5). The competition and mixing between these two components is computed in §6, with a net outcome depending on the mass range and on ambient conditions. This balance is of keen interest because interactions are likely to contribute, as noted above, key common features of apparently diverse astrophysical phenomena in the nearby and in the distant Universe. 2. DISORDER AND BRANCHING The language of Cayley trees with disorder (see Derrida & Spohn 1988) con- 6 veniently describes at the "microscopic" level of density contrasts δ how new col- lapses from the linear perturbations compete or combine with aggregations of the existing condensations. The tree is a computational structure (visualized by fig. 1) where at each step µ random weights wi are extracted in a cascade following a sequence of links, which may randomly branch into two. As the generation number µ increases, the progressive product of such random weights wi will be related to the probability of finding a fluctuation of the density field. The tree coordinate µ will be related to the mass scale, and the probability will be related to the number of condensations per unit mass. The end result at the "macroscopic" level will be a mass distribution N (M, t). The tree includes in one random process the following two components of the δ field: (1) disorder - with increasing resolution S ∝ σ2 ∝ M −2a the independent values taken at a given time by δ(M ) execute, as recalled in §1, a pure random walk; and (2) branching - δ may also jump by stochastic "branching", actually coalescing two paths into one. The statistics of the combined process is conveniently derived in terms of the partition function computed at each generation µ along the tree in the direction of coalescence from an initial µo: Zµ = P[paths] Πµ i=µo wi, where the product refers to the i-th preceding serial links of the tree, and the sum includes all paths coalesced at µ. The distribution function of Z will be Pµ(Z), with moments hZ kiµ ≡ Z dZ Pµ(Z) Z k . 7 (2.1) It proves technically convenient to set wi = 1 + vi, to exhibit the unbiased average wi = 1 when v fluctuates around 0. In this representation the partition function reads (1) Zµ = X[paths] Πµ i=µo (1 + vi) . (2.2) For the specific tree in fig. 1 the definition of Z implies the following recursion relations to hold at each elementary step dµ: Zµ+dµ = (cid:26) (1 + v) Zµ (1 + v)[Z1µ u1 + Z2µ u2] with probability 1 − ηdµ, with probability ηdµ. (2.3) The first line describes pure disorder, and the second includes branching; v and u are stochastic variables. Gaussian initial conditions for the perturbation field (1) For v → 0, the limit we shall consider, this is indistinguishable from the other vi . The latter is heuristically attractive because representation Zµ = P[paths] e one may directly identify Pi vi with Pi δi = δ and use the relationship δ = E/Eb with −Pµ i=µo the energy E of linear perturbations in a critical universe normalized to the background potential energy (Peebles 1980). The resulting Zµ = P[paths] e i=µo Ei/Eb has the −Pµ form of a standard partition function, and the counting expressed by the definition (2.2) implies a sum rule for the linear energies. Nonlinear superpositions of comparable en- ergy fluctuations are explicitly accounted for by P[paths] in the definition (2.2) and corresponding to the relation (2.3b). 8 imply for v a distribution which for small v goes into a Gaussian with variance D dµ proportional to the step length: g(v) = e−v2/2Ddµ/(2πDdµ)1/2 . (2.4) Another stochastic function r(u), generally far from symmetric, governs the distri- bution of the weights u translating, as we shall show, the interaction probabilities. A compact way to embody all moments is in terms of the generating function Gµ(x) = he−(1+x)Zµi . (2.5) Successive derivatives of G at x = 0 are related to moments of Z of increasing order, as shown by the formal expansion Gµ(x) = Xk (−1)k (1 + x)khZ kiµ/k! . (2.6) Correspondingly, the evolution of the moments can be embodied in a master differential equation equivalent to the recursion relations eq. 2.3. It follows from eq. 2.5 that in an interval dµ the two components simply add with the weights provided by their probabilities as given eqs. 2.3, and their superposition gives Gµ+dµ(x) = (1 − ηdµ) Gµ + η dµ G2 µ with (2.7) Gµ ≡ Z dv g(v) Gµ(x + v + vx), Gµ ≡ Z du r(u) Gµ(x u) . The continuous limit dµ → 0 yields ∂µGµ = D ∂xxGµ/2 + η( G2 µ − Gµ) . (2.8) 9 For, in this limit the lhs yields Gµ+dµ → Gµ + ∂µGµdµ, and on rhs the variance of v shrinks proportionally to dµ, so that Gµ → Gµ + [D + O(x)] dµ ∂xxGµ/2. When the two sides are set equal, finite terms cancel out, and to the first order in dµ the above equation obtains near x = 0, which will be the relevant point. It is easily perceived, and is discussed in detail in §3, that in the limit of no branching (i.e., η → 0) eq. 2.8 reduces to a pure diffusion equation for Gµ(x), similar to the equation given by Bond et al. 1991 and recalled in the Introduction. The opposite limit of branching with no Gaussian noise yields, as we shall see in §5, the Smoluchowski equation for the kinetics of N (M, t) under binary interac- tions discussed by CCM. We next substantiate these two limits by examining the relationship of Gµ(x) with N (M, t), and that of the generation number µ with the resolution S or with the physical time t. We stress that solving eq. 2.8 for Gµ(x) is equivalent to computing the single moments of Z directly from the relations 2.3 and then synthetizing Gµ(x) from its expansion 2.5. The advantage of following this latter route is that the recursion relations are very simple, and especially suited for numerical work. On the other hand, the master eq. 2.8 is convenient for making contact with previous work in limiting cases and for discussing the balance of the two competing modes. The statistical effects of this competition are illustrated in fig. 2. 3. FROM THE CAYLEY TREES TO THE MASS FUNCTION We first note that in the limit of no branching (i.e., η → 0) the remaining 10 terms of eq. 2.8 yield the structure of a diffusion equation ∂µGµ = D ∂xxGµ/2 . This, by the gauge (structure-preserving) transformation dµ = − 1 2 d ln S , dx2 ∝ dδ2 D/S , (3.1) (3.2) with the condition D dµ > 0 (see eq. 2.4), can be made identical with the equation (Bond et al. 1991, Lacey & Cole 1993) ∂SQ = ∂δδQ/2 , (3.3) which governs the evolution of Q(δ, S) dδ, the density of trajectories (random walks) of δ as the resolution S ∝ σ2 ∝ M −2a is increased or the scale is decreased. Because of its central role in what follows, we discuss the gauge 3.2 in more detail. We first integrate eq. 3.2b to x ∝ (δ − δc)/σ , (3.4a) so that x = 0 corresponds to δ = δc; thus the collapse threshold is embedded in the tree formalism as a zero point for the tree variable x. Then we specify (see fig. 3 and its caption) the relationship of the remaining tree coordinate µ and of its initial value µo with the physical variables M and t. At each time t the resolution scale S = M −2a corresponding to a given generation number µ is reckoned from a minimum So ∝ [Mc(t)/ǫ]−2a with ǫ ≪ 1, correspond- ing to a maximum mass Mmax ≡ Mc(t)/ǫ ≫ Mc(t), so that eq. 3.2a is integrated 11 in the form µ − µo = − 1 2 ln S So = ln (ǫ m)a . (3.4b) Note that So contains the t-dependence of the initial condition µo for each realiza- tion of the tree. The tree coordinates µo and µ− µo are used as independent ones in what follows. The mass scale M is an independent variable in the frame of the physical coordinates (the plane M, t in fig. 3), but in terms of the tree coordinates becomes a function of m and t through the relation M = m Mc(t) (see also caption to fig. 3). Having so specified the relations between the physical and the tree variables, we now elaborate a procedure for computing the mass function from the tree. Following the papers by Bond et al. 1991 and Lacey & Cole 1993, the PS selection rule in terms of Q(δ, S) and N (m) = N (M )dM/dm is written as N (m) m dm = − ρ Mc dZ δc −∞ dδ Q = − ρ Mc dS [∂δQ/2]δc . (3.5) The second equality, which expresses the mass fraction as a density in resolution of flux across the (absorbing) boundary δc, is formally provided by integration over δ of eq. 3.3. We note that within the above theory a Laplace transformation relates N (M ) to Q(δ, S). In fact, the integral form of the above relation may be recovered by applying the operator [∂δ]δc to both sides of the integral relation Z dm N (m)e−m(δ−δc) = ρ Mc Z dS Q/2 . (3.6) 12 In other words, the selection rule and the diffusion equation imply that R dS Q/2 is the Laplace transform of N (m) Mc/ρ. (2) Based on the close similarity of eqs. 3.1 and 3.3 when the gauge eqs. 3.2 are considered, we propose the following general form of the selection rule: N (m) NT m dm = −d Z dx Gµ(x) = −dµ [∂xGµ(x)]o = −dµhZµ e−Zµi. (3.7) Here NT (t) ≡ R N (m) dm ∝ ρ/Mc, and the derivative is computed at x = 0, corresponding to δ = δc after the gauge transformation 3.4a; note from the second term that additive components of G independent of µ will not matter. In integral form, the fraction of condensed mass is Z dm N (m) m = −NT Z dµ [∂xGµ]o , (3.8) where the last integral is actually t-independent (see footnote 2). But the above relation also obtains by differentiation ∂x at x = 0 of the relation Z dm N (m) e−m x = NT Z dµ Gµ(x) , (3.9) (2) In the following, unless otherwise specified, the integrals are meant to run over the full range of the variables. These are as follows: δ ∈ [−∞, +∞], m ∈ [Mmin, Mc/ǫ], and µ ∈ [−∞, µo]. The integrals over µ ranging in the last interval may be rewritten as integrals over µ − µo in the corresponding range, and are explicitly t-independent. 13 which expresses the generating function, normalized to NT , as a Laplace transform of the mass distribution. The latter then obtains by anti-transforming Gµ(x). Equivalently, all successive moments of N (m) are obtained by successive differentiation of eq. 3.9. For example, the zeroth-order moment is given by R dmN (m) = NT R dµGµ(0), which fixes the normalization R dµGµ(0) = 1; the 1st moment is given by eq. 3.8, consistent with the differential form 3.7. 4. DISORDER: THE PS LIMIT We now show how the PS mass distribution may be derived from eq. 3.7 in the limit of no branching, that is, η → 0. The latter equation involves [∂xGµ]o = h Zµ e− Zµi = Xk=1 (−1)k−1hZ kiµ/(k − 1)! . (4.1) The single moments of Z will be directly derived from the recursion relation eq. 2.3a (with D = 1) in the form hZ kiµ+dµ = Z dv e−v2/2dµ (1 + v)k √2πdµ hZ kiµ . (4.2) Expanding the binomial around v = 0 to the lowest (second) significant order, integration of eq. 4.2 yields hZ kiµ+dµ = h1 + k2 − k 2 dµihZ kiµ . In the continuum limit dµ → 0 this yields d lnhZ kiµ = k2 − k 2 dµ , 14 (4.3) (4.4) which using the gauge 3.2a (i.e., dµ = −d lnσ) integrates to hZ kiµ/hZ kio = (σo/σ) k 2 −k 2 . (4.5) This relation implies that the first moment of Z is a constant relative to µ, which is a natural consequence of the property w = 1. The next two moments yield the leading contributions to the sum 4.1. In fact, following eq. 4.5 and keeping only the µ-dependent terms (as noted just after eq. 3.7), we find h Zµ e− Zµi = hZ h1 − Z + Z 2 2 + ...ii = Z 2 o σo σ h1 − Zo 2 σ2 o σ2 + O(cid:16) σ5 σ5(cid:17)i . o (4.6) Recalling from §3 that σo/σ = [M/Mmax]a = (ǫ m)a, it is seen that σo/σ < 1 holds for M < Mmax = Mc/ǫ, and the last equality can be resummed to within O(ǫ5a) to yield [∂xGµ]o ≈ Z 2 o σo σ e−Zo σ2 o /2 σ2 = Z 2 o (ǫ m)a e−Zo(ǫ m)2a . (4.7) The tree by itself, as any statistics, does not specify the dynamics of gravi- tational collapses, and provides only scaling behaviors: these, on substituting eq. 4.7 in eq. 3.7, are N (m) = ρ Mc m−2 d ln σ d ln m [∂xGµ]o ≈ 2 a ρ Z 2 o Mc (ǫ m)−2+a e−Zo (ǫ m)2a/2 , (4.8) and turn out to be the same as in the PS distribution. As to the constants which do carry dynamical information, the initial condition Zo may be set to δc, the counterpart here of the boundary condition used by Bond et al. 1991. In addition, the rescaling N (ǫ m) = N (m)/ǫ holds, and the overall normalization follows from 15 the requirement R dµ Gµ(0) = 1 derived in §3. Thus the full PS expression 1.1 is recovered. 5. BRANCHING: THE SMOLUCHOWSKI LIMIT We now consider eq. 2.8 in the opposite limit of branching with no Gaussian noise. Then the remaining terms on rhs read ∂µGµ = η( G2 µ − Gµ). (5.1) Equivalently, the change of single moments of Z may be derived directly from the recursion relation 2.3b with no noise, to yield hZmiµ+dµ = dµ Z Z dZ1 dZ2 P (Z1) P (Z2) (Z1µ + Z2µ)m +(1 − dµ)Z dZ P (Z) Zm µ ; (5.2) here the coupling parameter η has been absorbed into a rescaling of µ for con- venience as will become apparent after eq. 5.11. Expanding the binomial, this becomes hZmiµ+dµ = dµZ Z dZ1d Z2 P (Z1) P (Z2) m Xk=1 m! k! (m − k)! 1µ Zm−k Z k 2µ +(1 − dµ)hZmiµ. In the continuous limit dµ → 0 this yields ∂µhZmiµ m! = m Xk=1 hZ kiµ k! (m − k)! − hZmiµ hZm−kiµ m! . 16 (5.3) (5.4) This relation has a structure similar to the Smoluchowski equation which governs the kinetics of binary aggregations of definite condensations (CCM): ∂N ∂t = dM ′ K(M ′, M − M ′, t) N (M ′, t) N (M − M ′, t) 0 1 2 Z M −N (M, t)Z ∞ 0 dM ′ K(M, M ′, t)N (M ′, t) . (5.5) The kernel K represents the rate of aggregation in a system of condensations with relative velocity V and interaction cross section Σ, and averages to τ −1 = NT ΣV in terms of the component number density NT (t). Indeed, we show next that the above equation may be identified with eq. 5.4 (or with the equivalent eq. 5.1) when dealing with high-contrast condensations insensitive to the perturbation field; that is, in the limit of no disorder (with δ = σ) and of high-contrast condensations (i.e., δc → 0). For a constant kernel K we again proceed directly. To this end, we first rewrite the discretized form of eq. 5.5, after dividing by NT /2, in the form 2 N 2 T ∂NM ∂t + NM NT = XM ′ NM −M ′ NT NM ′ NT − NM NT , (5.6) where the parameter K has been absorbed into a rescaling of t for convenience which again will become apparent after eq. 5.11 below. We then express the t-derivative in terms of the tree variable µ. The gauge eq. 3.2a yields dµ = − 1 2 d lnS = a d ln M . In the tree frame (as we discussed in §3) M = m Mc(t) ∝ m/NT hold, and one obtains dµ ∝ − dNT NT = 1 2 NT dt , 17 (5.7) with the last equality coming from eq. 5.5 integrated over M . Then eq. 5.6 can be written as ∂ ∂µ NM NT = XM ′ NM −M ′ NT NM ′ NT − NM NT . (5.8) Now the above equation exhibits even more clearly a structure similar to eq. 5.4, and is identical to it provided that NM NT ∝ hZmiµ m! (5.9) holds. The simplest way to prove this relationship is to to write the integral relation 3.9 with the shorthand 1 + x = w in the form Z N (m) NT wm dm = Z dµhe−wZµi = Xk wk Z dµ hZ kiµ k! , (5.10) with the signs (−1)k included into hZ ki in view of the invariance of eq. 5.4 relative to such transformations. Then we note that eq. 5.4 for m = 1 implies d lnhZiµ ∝ dµ, and hence hZiµ ∝ NT by eq. 5.7. For m = 2 it implies hZ 2i ∝ NT asymptotically when NT ≪ NT o holds; similarly for the higher orders. With hZ kiµ/NT k! →const relative to µ, the integration on rhs of eq. 5.10 reduces to R dµNT = R dµe−µ in view of eq. 5.7, and yields a constant factor. Thus, writing the integral on the lhs of eq. 5.10 in discrete form, we find Xm Nm wm ∝ Xk wkhZ kiµ/k! , (5.11) where of course the dummy indexes k and m may be identified. We now apply to both members the operator [∂m w ]w=0, which in fact is appropriate for high- contrast structures with δc → 0 and δ ≈ σ. The result is the relation 5.9, thus 18 completing the identification of the Smoluchowski equation of the form 5.5 with the tree recursion relations of the form 5.4. The proportionality constant in eq. 5.11 translates into a rescaling by a constant factor of the correspondence between the independent variable in eq. 5.4 (which in full is µη), and that in eq. 5.6 (which in full is t/τ ). The statement that PmhZmiµ/NT m! becomes time-independent holds for the solutions N (M, t) of the Smoluchowski equation in their asymptotic, self-similar stage where N (M )/N 2 T is time-independent. The equivalence of the Smoluchowski equation 5.5 with a Cayley tree can also be proved when the kernel K(M, M ′) is mass-dependent. Then the equivalence may be recovered by going through the Laplace transform of eq. 5.5, as has been carried out explicitly by Peshanski (1992) for the case of multiplicative kernels K(M, M ′) = K(M ) K(M ′). The link between kernel and weight distribution on the tree r(u) is explicitly given by K(m) = C Z ∞ 0 du r(u) eum , (5.12) which amounts again to a Laplace transformation. The constant C ∝ τ 1/2 contains the normalization of K, since r(u) is normalized to 1. 6. DISORDER AND BRANCHING SUPERPOSED We now compute numerically the mass function in conditions where both disorder and branching are effective. As has been said, for numerical work it is much easier and faster to use directly the recursion relations (eq. 2.3) for the 19 partition function Zµ – rather than the master equation 2.8. This is done by the following procedure. At each step µ of the tree a numer- ical algorithm extracts the weights v from a Gaussian distribution, and uses the first recursion relation 2.3 with probability 1 − η ∆µ, or the second with proba- bility η ∆µ, to construct the partition function at the next step. In the present computations we use a delta-function r(u) = δ(u), corresponding after eq. 5.12 to a constant interaction kernel. In fig. 2 we have shown the resulting trajectories of the contrast for increasing S, which are related to the tree variables x and µ by the gauge eqs. 3.4. For each tree, we generate the entire Zµ and its moments. The procedure is repeated for a large number of trees (up to 103). The moments of the partition function are used to compute [∂xGµ]o = Xk=1 (−1)k−1hZ k µi/(k − 1)!, (6.1) where the sum we actually used runs up to k = 10. Then at any given t we compute the mass distribution N (M, t) following eq. 3.7, and check that the total mass is conserved in time to within 1% when 103 trees are used. The result depends on the branching probability η, which acts like an effective coupling constant for binary interactions. For small η ∼< 10−2 the diffusion part of eq. 2.8 dominates and one recovers the PS mass distribution. For increasing values of η the resulting distributions become steeper, while the cutoff at large masses tends to straighten up. When η ∼ 10−1 the distribution reaches its asymptotic form, a pure power law with logarithmic slope close to −2; see fig. 4 20 There are two reasons why even a small value of η is effective. First, note that on transforming eq. 2.8 following the gauge eqs. 3.4 the effective coupling parameter is η/S ∝ ηm2a, that is, a coupling more effective at larger masses. Second, a kernel constant in time, as considered here, maximizes the interactions because it implies no dependence of the number density on ambient evolution. This is physically realistic for interactions within large scale structures surviving for longer then τ . These results compare interestingly with the results from cosmological N-body simulations with a large dynamical range and highly resolved data analysis, like those performed by Brainerd & Villumsen 1992 and Jain & Bertschinger 1994. These papers agree in finding a slower evolution and an excess at large masses compared with the PS mass distribution, with the former finding consistency with a simple power law distribution, approximatively M −2 within the walls and fila- ments. These features are consistent with our findings. In terms of overdensities vs. resolution as used in fig. 2, the interactions cause not only an obvious density decrement of the trajectory distribution toward larger coalesced masses, but also a skewness relative to the Gaussian counterpart. 7. DISCUSSION The existing theories for the shape and the evolution of the mass distribution N (M, t) fail to capture the full complexity of cosmic structures. The PS the- ory, with its recent elaborations, provides with eq. 1.1 the best quantification for 21 the scenario envisaging hierarchical collapses from initial density perturbations. Yet the result differs as to amplitude, shape, and evolution from data or from simulations and from realistic excursion set computations with the same input parameters. The alternative mode to hierarchically building up structure is based on binary aggregations between high-contrast condensations. This predicts non- Gaussian formation of rare large objects, and describes more closely the erasure of substructures within a structure by resolving timescales different from its dy- namical time td. But it requires input, at least initially, of formed condensations into an environment protected from the Hubble expansion (CCM). We submit that both these modes constitute only partial representations of the evolution of the density field. They select either purely Gaussian random walks of the linear density contrasts as functions of the resolution, up to crossing the threshold of nonlinearity; or trajectories "branching" stochastically only at large values of the contrast. Correspondingly, the PS function 1.1 and the Smoluchowski aggregation equation 5.5 correspond to restricted, "macroscopic" averages from a complete " microscopic" field statistics which treats aggregations and direct collapses as proceeding together at similar contrast levels. We propose such a complete statistics which combines, in the form of a Cayley tree (or random cascade) as represented in fig. 1, random walk and stochastic branching of fluctuations into one partition function governed by the single master equation 2.8. We also propose the Laplace transform relationship eq. 3.8 between the tree generating function Gµ(x) and the mass distribution N (M, t). 22 We have proved that our proposals indeed yield in the appropriate limits both the diffusion equation (§4) with the associated PS function eq. 1.1, and the Smoluchowski equation (§5). The former limit applies to Gaussian fluctuations that independently reach the nominal threshold for collapse and virialization. The latter limit is derived with no formal recourse to a threshold, and describes binary interactions of an initially given set of already formed condensations. Pleasingly, from averaging over the tree in the latter limit one obtains a mean-field, kinetic equation with different solutions depending on the kernel, and evolving away from the initial form; in the former limit, one instead obtains a single PS function. Both limiting processes are hierarchical and imply "merging", i.e., inclusion of smaller condensations into larger ones. But pure DHC in fact envisages only reshuffling of (generally) many condensations which belong to lower hierarchical levels into a higher level on a larger mass scale at a subsequent time, strictly following the conditions set ab initio in the linear fluctuation field. The natural quantification of this process is provided by the conditional probability that a trajectory up-crossing the threshold at the resolution S2 had a previous up-crossing at S1 > S2 (see Bond et al 1991; Bower 1991; Lacey & Cole 1993). Pure aggregation, on the other hand, envisages pairs of condensations coa- lescing into a third, at a similar – and high – contrast level. Here the nonlinear interactions are stochastically set by ambient density and relative velocities, and may be triggered at any time when larger scale structures outline volumes with expansion slowed down relative to the general "field". 23 The Cayley tree approach not only proves successful in deriving the two, op- positely extreme modes for hierarchically building up structures, but also provides a number of links between them: the common tree algorithm or the master equa- tion 2.8; the Laplace transform relationship equation 3.8 of Gµ(x) with N (M, t), which extends the PS selection rule to interacting fluctuations; and the tree vari- able µ that includes both the resolution S (used to derive the PS function) and the physical time t (used to derive the Smoluchowski equation), as stressed below. In fact, we have seen that dµ = − 1 2 d lnS holds, appropriate to the former case. On the other hand, from the relation M = m Mc(t) ∝ m/NT one obtains ∂tµ ∝ −d lnNT /dt, which is the eq. 5.7 used in deriving the Smoluchowski equa- tion. The constant factors (and specifically the spectral index n, the last remnant of the initial perturbation spectrum) are actually irrelevant in the subsequent identification of eq. 5.5 with the tree recursion relation in the absence of Gaussian noise, as expected. Above all, the tree formalism describes with equal ease the mixing and com- petition of the two pure modes; that is, it describes condensations interacting and aggregating while still growing or collapsing. The process may be computed ei- ther from the master equation, or directly in terms of the tree partition function, which is very convenient and fast on computers. These computations yield, when branching is fully effective, N (M, t) in the form of a steep, slowly lowering power law with logarithmic slope ≈ −2. The result is due to the continuous input from the Gaussian noise over many scales, combined with the increasing effect of the 24 branching mode at larger masses. In fact, even a value of η < 1 is effective because the relevant coupling param- eter η/S ∝ η m2a increases at higher masses. Actually, the interactions are maxi- mized by a kernel constant in time as considered here, corresponding to densities and velocities unaffected by expansion. This is physically realistic for interactions taking place within large-scale structures which survive for times longer that τ . In these conditions high-mass condensations form faster than the Gaussian rate within large scale structures, not unlike the formation action seen in large N-body simulations (Brainerd & Villumsen 1992; Babul & Katz 1992). Observations show that groups are concentrated in filaments and sheets outlined by redshift surveys of galaxies (Ramella et al. 1990) and that the mass distribution over scales from 1013 M⊙ to several 1015 M⊙ is consistent with a steep power law (Giuricin et al. 1993). We shall discuss elsewhere the relationship of our approach with the dynamical descriptions of matter field under gravity – e.g., the adhesion model (see Shandarin & Zel'dovich 1989) or the frozen-flow approximation (Matarrese et al. 1992) – which include shear effects in addition to interactions of comparable fluctuations. Conditions of aggregating interactions protected from the Hubble expansion within larger structures are likely to underlie apparently diverse astrophysical phenomena, from formation of one large cD-like galaxy by strong interactions of members in groups, to the very formation and growth of groups and clusters within large scale filaments and sheets. Toward such a complex emergence of cosmic structure Cayley 25 trees provide a unifying approach. Acknowledgments. It is a pleasure to thank Margaret Geller for stimulating ex- changes, and Felix Ritort for helpful discussions. We acknowledge financial support by ASI, MURST and the EC HCM Programme. 26 REFERENCES Abell, G.O. 1958, ApJS., 3, 211 Babul, A., & Katz, N. 1992, ApJ, 406, L51 Bahcall, N.A., & Chokshi, A. 1991, ApJ, 380, L9 Barnes, J.E., & Hernquist, L.E. 1991, ApJ 370, L65 Bhavsar, S.P. 1989, ApJ, 338, 718 Bond, J.R., Cole, S., Efstathiou, G., & Kaiser, N. 1991, ApJ, 379, 440 Bower, R.J. 1991, MNRAS, 248, 332 Brainerd, T.G., & Villumsen, J.V. 1992, ApJ, 394, 409 Cavaliere, A., Colafrancesco, S., & Menci, N. 1992, ApJ, 392, 41 (CCM) Cavaliere, A., Colafrancesco, S., & Scaramella, R. 1991, ApJ, 380, 15 Derrida, B., & Spohn, H. 1988, J. Stat. Phys., 51, 817 Giuricin, G., Mardirossian, F., Mezzetti, M., Persic, M., & Salucci, P. 1993, SISSA preprint Gunn, J.E., & Gott, J.R. 1972, ApJ, 176, 1 Heckman, T.M. 1993, STScI preprint Jain, B. & Bertschinger, E. 1994, MIT preprint Jones, J.C. & Forman, W. 1992, in Proc. NATO-ASI Clusters and Superclusters of Galaxies, ed. A.C. Fabian (Dordrecht: Kluwer), p. 49 Katz, N., Quinn, T. & Gelb, J.M. 1993, MNRAS, 265, 689 Kolb, E.W., & Turner, M.S. 1990, The Early Universe (Redwood City, CA: Addison-Wesley) 27 Lacey, C., & Cole, S. 1993, MNRAS, 262, 627 Matarrese, S., Lucchin, F., Moscardini, L., & Saez, D. 1992, MNRAS, 259, 437 Peebles, P.J.E. 1965, ApJ, 142, 1317 --. 1980, The Large Scale Structure of the Universe (Princeton: Princeton Univ. Press) --. 1993, Principles of Physical Cosmology (Princeton: Princeton Univ. Press) Peshanski, R. 1992, CERN preprint TH.6044/91 Press, W.H., & Schechter, P. 1974, ApJ, 187, 425 Ramella, M., Geller, M.J. & Huchra, J.P. 1990, ApJ, 353, 51 Rees, M.J, Ostriker, J.P. 1977, MNRAS, 196, 381 Shandarin, S.F & Zel'dovich, Ya.B. 1989, Rev. Mod. Phys., 61, 185 Schechter, P.L. 1976, ApJ, 203, 297 Shlosman, I. 1990, in IAU Colloq. 124, Paired and Interacting Galaxies, ed. J. Sulentic & W. Keel (Dordrecht: Kluwer), p. 689 von Smoluchowski, M. 1916, Phys. Z., 17, 557 White, S.D.M., Briel, U.G., & Henry, J.P. 1993, A&A, 259, L31 28 FIGURE CAPTIONS Fig. 1. A schematic representation of a Cayley tree. The random process is represented as a cascade proceeding from the bottom up. At each step (labeled by the generation number µ + dµ) a random weight w = 1 + v is extracted according to a Gaussian distribution g(v). In addition, the path may either result from coalescing (with probability η dµ) of two branches that were still distinct at µ, or proceed along a single branch. The partition function at a point µ + dµ is constructed by summing the products of the weights over all paths leading to it. The heavy line marks a specific path on the tree with weights w1, w2, ... w7. Fig. 2. Contrast vs. resolution (in arbitrary units) of a number of trajectories computed from numerical realizations of the tree (details are given in §6). Top panel: Pure Gaussian noise resulting in random walks with dispersion increasing with increasing resolution (or decreasing mass). Bottom panel: Effects of including random branching (with η = 0.1) in the statistics of the trajectories; the branching points are marked by a dot. The study of the apparent differences between the two panels constitutes the thrust of this paper. Fig. 3. Relations between the physical variables M and t and the tree coordinates µ and µo . At a given t (top panel), the initial value So for the resolution corre- sponds to a maximum mass scale Mmax = Mc(t)/ǫ ≫ Mc(t). It also corresponds 29 to a value of the tree generation number µo ∝ lnMmax following eq. 3.2. So µ− µo ∝ ln m holds, with m ∝ M/Mc(t). For a given mass M (bottom panel), dif- ferent values of t correspond to various starting points µo, and to different values of µ visualized by the intercepts of a horizontal line for increasing t. Fig. 4. Mass function N (M ) resulting from numerical realizations of trees with both branching and disorder. The partition function and the related moments are computed from the eq. 2.3. Dotted curve: η = 0; dashed curve: η = 0.05; solid curve: η = 0.1. The first is identical with the Press & Schechter function, eq. 1.1. 30
astro-ph/0012006
1
0012
2000-12-01T02:45:36
Analytical Studies on the Structure and Emission of the SS433 Jets
[ "astro-ph" ]
We study the structure and emission of the SS 433 jets in the X-ray emitting region and in the inner and hotter portion inside the X-ray emitting region. In order to consider the jet structure from the inner to outer regions we develop the hybrid model combining the conical beam and the model beam whose cross section grows with the distance more slowly. We find that the jet beams in the inner and hotter portion are of two-temperature and emit a large amount of high energy gamma photons. Our analyses suggest the thick absorbing envelope to exist in the SS 433 system. Based on our results, we discuss the possible acceleration mechanism for the SS 433 jets.
astro-ph
astro-ph
PASJ: Publ. Astron. Soc. Japan , -- 12 (2018) Analytical Studies on the Structure and Emission of the SS 433 Jets Hajime INOUE, 1 Noriaki SHIBAZAKI, 2 and Reiun HOSHI2 1 The Institute of Space and Astronautical Science, 3-1-1 Yoshinodai, Sagamihara, Kanagawa 229-8510 E-mail: [email protected] 2 Department of Physics, Rikkyo University, 3-34-1 Nishi-Ikebukuro, Tokyo 171-8501 (Received ; accepted ) Abstract We study the structure and emission of the SS 433 jets in the X-ray emitting region and in the inner and hotter portion inside the X-ray emitting region. In order to consider the jet structure from the inner to outer regions we develop the hybrid model combining the conical beam and the model beam whose cross section grows with the distance more slowly. We find that the jet beams in the inner and hotter portion are of two-temperature and emit a large amount of high energy gamma photons. Our analyses suggest the thick absorbing envelope to exist in the SS 433 system. Based on our results, we discuss the possible acceleration mechanism for the SS 433 jets. Key words: Astrophysical jets -- Stars: individual (SS 433) -- X-rays: binaries 0 0 0 2 c e D 1 1 v 6 0 0 2 1 0 0 / h p - o r t s a : v i X r a (Accepted for publication by Publ. Astron. Soc. Japan) 1. Introduction The galactic source SS 433 is well-known as an unique object exhibiting a jet phenomenon ( for review, see Margon 1984 and Vermeulen 1989). The moving behavior of Balmer and He I lines has been interpreted successfully in terms of the two opposing jets precessing with a period of 164 days. The velocity of matter outflowing in the jet is found 2 Author(s) [Vol. , to be 0.26c and remarkably stable, where c is the speed of light. The SS 433 system is an eclipsing binary with the orbital period of 13 days. The system may consist of an early type star and an accretion disk around a compact object. The compact star may be a neutron star or a black hole, although the definite determination has been still waited. SS 433 has attracted huge attention from the time of the discovery because of its peculiar behavior. A lot of observations have been conducted in the radio, optical and X-ray bands. Many theoretical studies have also been put forward. In spite of these enormous efforts the most fundamental problems on SS 433 still remain unsolved. Those are the nature of the compact star and the acceleration and collimation of the jet beams. Recently, the ASCA satellite has detected X-ray lines of various elements, such as Fe, Ni, Mg, Si, S and Ar, from the SS 433 jets ( Kotani et al. 1996; Kotani et al. 1997). In the Fe case both the blue-shifted and red-shifted X-ray lines, which are emitted respectively from the approaching and receding jets, have been observed at the same time. Fitting the analytical jet model to these ASCA data, Kotani et al. (1996, 1997) have examined the properties of the X-ray emitting portion of the jet beams. They have derived the important constraints on the jet parameters such as the density, temperature and size of the X-ray emitting region and the mass outflow rate. Their analyses have also implied the X-ray absorbing gas to exist in the system. We study the structure and emission of jet beams in the X-ray emitting region and in the inner and hotter portion inside the X-ray emitting region. Based on the results on the jet structure, we discuss the acceleration mechanism for the SS 433 jets. We present a simplified model for the SS 433 jets in section 2. We derive the analytic solutions for the structure and emission of the X-ray jets in section 3. Deriving the analytic solutions, we show that the jet structure is of two-temperature in the inner and hotter portion inside the X-ray emitting region in section 4. In the final section we discuss the appropriate jet model, the acceleration mechanism and the absorbing envelope based on our analytical results on the jet structure and emission. 2. Model The jet may be divided into several characteristic regions depending on the distance from the central engine, the inner region, the X-ray emitting region, the optical and radio emission regions, and the outer region. Here, we focus on the X-ray emitting region and the hotter region inside of it. No. ] Running Head 3 No observational evidence is found for the velocity variation along the jets in SS 433. This fact implies that the matter entrainment from the ambient medium and other braking processes are unimportant in the region of our interest. Hence, we assume the matter velocity constant, v = 0.26c, along the jets. The mass outflow rate M in a jet is written as M = Sρv , (1) where r is the distance from the central engine, ρ the density and S the area of the jet cross section. For simplicity, we consider the beam shapes as expressed by S = πR2 ∝ rn , (2) where R is the radius of the jet cross section and n the constant number (Fukue 1987). In the following we examine the jet structure and emission especially for the cases of n = 1 and n = 2. The case of n = 2 corresponds to the conical beam, while the case of n = 1 to the jet beam whose cross section grows with the distance more slowly than that of the conical beam. The matter in the jets cools, expanding and emitting radiations. Writing the temperature and radiative loss rate respectively as T and Λ, we have for the energy equation 3 2 k µmH dT dr − kT µmH 1 ρ dρ dr = − Λ v , (3) where mH is the mass of hydrogen, k the Boltzmann constant and µ the mean molecular weight of matter. The second term on the left hand side of equation (3) and the term on the right hand side represent the adiabatic and radiative coolings, respectively. We normalize physical quantities with those at a fiducial point r0 ; x = r/r0, y = T /T0, z = ρ/ρ0 and f = Λ/Λ0. Here and in the following the subscript 0 is used to denote the physical quantities at r = r0. Using these parameters, equations (1) and (3) reduce to the dimensionless equations, z = x−n and dy dx + 2 3 n 1 x y = −ξ0f , (4) (5) 4 Author(s) [Vol. , respectively. In equation (5) the free parameter ξ0 is the ratio of the flow time tf 0 to the radiative cooling time tr0, ξ0 = tf 0 tr0 , where tf 0 and tr0 are defined respectively by tf 0 = r0 v and tr0 = 3 2 kT0 µmH Λ0 . (6) (7) (8) Note that the flow time also expresses the adiabatic cooling time due to the plasma expansion. Equation (5) shows that the free parameter ξ0 determines the property of solutions and hence the thermal structure of the jet. 3. One-Temperature Jets We consider here the region of the jet beams, where a typical temperature is ∼ 10 keV and X-rays are emitted. 1, where τe is the optical thickness We assume that the jet is optically thin to X-ray photons, satisfying τe = κeρR < ∼ and κe is the electron scattering opacity. In this region the ion-electron scattering occurs so frequently that the electrons and ions have the same temperature. 3.1. Analytic Solutions When the radiative cooling rate is given by the thermal Bremsstrahlung emission, Λ = Λff = 5.7 × 1020ρ√T ergs g−1s−1 , the normalized cooling rate is obtained as f = z√y . Equation (5) together with equations (4) and (10) can be solved analytically to yield √y =   1 x1/3 + 3 2 ξ0 (cid:0) 1 x1/3 − 1(cid:1) 1 x2/3 + 3 2 ξ0 (cid:0) 1 x − 1 x2/3(cid:1) for n = 1 for n = 2 . (9) (10) (11) No. ] Running Head 5 Note that the solutions (11) tend to the adiabatic cooling formula, y = 1/x2/3 and y = 1/x4/3 respectively for n = 1 and n = 2, as ξ0 approaches to zero. If we intergrate the Bremsstrahlung emission along the jet, we have for the X-ray luminosity emitted from the region outside the distance x, L(x) = S0r0ρ0Λff 0 g(x) ergs s−1 , where g(x) =   3(cid:0)1 + 3 2 ξ0(cid:1)(cid:16) 1 x1/3 − 1 max(cid:17) − 3 1/3 x 2 ξ0 ln(cid:0) xmax x (cid:1) for n = 1 3 5 (cid:0)1 − 3 2 ξ0(cid:1)(cid:16) 1 x5/3 − 1 x max(cid:17) for n = 2 max(cid:17) + 3 4 ξ0 (cid:16) 1 5/3 x2 − 1 x2 (12) (13) . In equation (13) xmax denotes the outer end of X-ray emitting region. As an xmax we choose here the distance beyond which the temperature drops below ∼ 0.1 keV. Adopting the solution for the conical beam (n = 2) at r/r0 ≥ 1, Kotani et al. (1996) have calculated the line X-ray emission from Fe, Ni and other elements and made the detailed comparison with the ASCA observations for SS 433. We list the standard set of jet parameters they obtained in Table 1. In Table 1 we also list the jet parameters derived from using the solution for the model with n = 1. The solution of n = 1 also reproduces well the observed properties, such as the X-ray luminosity ( ∼ 1036 ergs s−1) and temperature ( ∼ 20 keV). In both solutions the fiducial point is chosen to be the base of the X-ray emitting region of the jet. The length of the X-ray emitting region is obtained as ℓX ∼ 1013 cm. The optical thickness across the beam is τe0 ∼ 0.1 and hence the jet plasma in the X-ray emitting region is transparent, justifying the assumption made in the energy equation. Note that the value of the dimensionless parameter ξ0 is ∼ 0.1 or less. This fact shows that the temperature structure is determined mainly by the adiabatic cooling loss due to the plasma expansion. We notice that the kinetic energy LK carried by the outflowing matter amounts to more than 1039 ergs s−1 and exceeds the Eddington luminosity for a star of one solar mass. 3.1.1. Jet inside the X-Ray Emitting Region The jet radius R0 of the X-ray emitting region and its distance r0 from the central engine are at least several orders of magnitude larger than the size of the central engine that drives the jet, a neutron star or a black hole. Although Kotani et al. (1996, 1997) have truncated the inner portion of the jet, it is quite natural to assume that 6 Author(s) [Vol. , the jet beam extends more down to the vicinity of the central engine. In the following let us examine the property of the jet expected to exist inside the X-ray emitting region. We apply the solution (11) to the inner region adjacent to the X-ray emitting region. Following equations (4) and (11), the density and temperature of the jet increase with decreasing distance. Hard X and soft gamma rays are emitted from the jet. The energy transfer from ions to electrons through the coulomb scattering becomes less efficient at higher temperatures, whereas electrons lose energy promptly radiating X and gamma rays and expanding adiabatically. The insufficient energy transfer makes the ion and electron temperatures different at T > ∼ 109 K (see the next section for detail). The jet reaches the critical temperature T1 ∼ 109 K at the distance r1. Here and in the following we denote the physical parameters at the innermost region of the one-temperature jet by the subscript 1. We list the physical quantities at r = r1 in Table 2. The relativistic Bremsstrahlung emission rate and adiabatic cooling rate, which are mentioned in the next section, are used to calculate ξ1 in Table 2. The optical thickness across the beam is less than one, although approaches one, and the jet is still optically thin against electron scattering near r = r1. Note that the parameter ξ1 in the case of n = 2 (conical case) exceeds, though a little, one. The radiative loss as well as the adiabatic loss also contribute significantly to the plasma cooling near the innermost region of the one-temperature jet in the conical case. In the n = 1 case the cooling of the jet plasma is still determined by the adiabatic loss. The luminosity of X and soft gamma rays emitted from the region at r ≥ r1 is calculated from equations (12) and (13) and is listed in Table 2. In the n =2 case the luminosity of soft gamma rays emitted from the region between the distances r1 and r0 exceeds the luminosity of X-rays (∼ 1036 ergs s−1) emitted from the region outside the distance r0 approximately by an order of magnitude, while in the n = 1 case the both luminosities are comparable. 4. Two-Temperature Jets We have the two-temperature jet inside the distance r1, where the ion temperature is higher than the electron temperature. Ions transfer energy to electrons through the ion-electron scattering, which in turn lose energy by the radiative emission and adiabatic expansion. The distribution of ion temperature along the jet beams can also be described by equation (3), taking µ = 1 and substituting the ion energy loss rate Λie for the radiative loss rate Λ. No. ] Running Head The ion energy loss rate Λie is given approximately by Λ = Λie = 3 2 kT mH νE , 7 (14) where T now expresses the ion temperature and νE is the ion-electron collision frequency. The ion-electron collision frequency is written as νE = 3.6 × 1022 ρ T 3/2 e s−1 , where Te is the electron temperature. (15) We adopt the relativistic Bremsstrahlung emission with the Compton amplification for the radiative energy loss of electrons and approximate the emissivity by Λff = 4.9 × 1016ηρTe ergs g−1s−1 , (16) where η is the Compton amplification factor. Equation (16) can be used approximately even in the optically thick case if the photon diffusion time across the jet is shorter than the flow time. Energy loss rate of electrons due to the adiabatic expansion can be approximated by Λad = kTe mH 1 tf . (17) The electron temperature is determined from the balance between the energy acquired from ions and the energy lost to radiation and adiabatic expansion. When the radiative loss is dominant over the expansion loss (Λff ≫ Λad), the electron temperature is expressed as Te = 3.8 × 105η−2/5T 2/5 K , whereas when Λff ≪ Λad, Te = 2.0 × 105T 2/5 K . (18) (19) In equation (19) the physical parameters for the n = 1 case in Table 1 are used. If we equate the ion and electron temperatures in equations (18) and (19), we can derive the critical temperature T1 above which the jet plasma is in the two-temperature regime, T1 ∼ 2 × 109η−2/3 K for Λff ≫ Λad (20) 8 and Author(s) T1 ∼ 6.9 × 108 K for Λff ≪ Λad . [Vol. , (21) We choose the position of r ∼ r1 as the new fiducial point and normalize the physical quantities with those at r = r1 ; x = r/r1 , y = T /T1 , and z = ρ/ρ1. The dimensionless equation (5) with the free parameter ξ1 replaced for ξ0 is again derived from the ion energy equation. The parameter ξ1 is now re-defined as the ratio of the flow time to the ion energy loss time or the ion-electron scattering time given by tie = ν−1 E . Furthermore, the ion energy loss rate is written in the dimensionless form as f = zy2/5 . We solve equation (5) together with equations (4) and (23) and derive the analytic solution, y3/5 =   1 x2/5 + 3 2 ξ1 (cid:0) 1 x2/5 − 1(cid:1) 1 x4/5 + 3ξ1 (cid:0) 1 x − 1 x4/5(cid:1) for n = 1 for n = 2 . (22) (23) (24) Using equation (16) and integrating the high energy emission along the jet, we obtain the luminosity of high energy photons emitted outside the distance x. The luminosity L(x) is approximately given by L(x) ∼ S1r1ρ1Λff 1h(x) ergs s−1 , where h(x) =   15 4 (cid:0)1 + 3 2 ξ1(cid:1)2/3 1 x4/15 for n = 1 3 5 1 x5/3 for n = 2 . (25) (26) The outer solutions (11) are connected to the inner solutions (24) at r = r1, thereby yielding a break in the temperature distribution. As we proceed inwards along the jet, the ion and electron temperatures increase follow- ing equations (18), (19) and (24). Moreover, the ion temperature increases faster than the electron temperature, rendering the jet of two-temperature. As we move furthermore inwards, we reach the sonic point where the ion No. ] Running Head 9 sound velocity becomes equal to the flow velocity. Although derived by ignoring the pressure gradient force and by assuming the constant flow velocity, the above solution may be used down to the sonic point as long as the approximate estimates of the jet parameters are concerned. We list the physical quantities at the sonic point, which are denoted with the subscript s, in the Table 3. Here, for simplicity, the compton enhancement factor is set as η ∼ 1. In the conical jet case electrons lose energy mainly by emitting radiations. In Table 3 the radiative cooling is adopted for the n = 2 case. The sonic point is located very far from the central engine. The ion temperature is approximately an order of magnitude higher than the electron temperature. The thermal equilibrium is not complete among ions since the ion-ion scattering time exceeds the flow time. As seen from the values of ξ1 and ξs, the radiative cooling as well as the adiabatic cooling contributes significantly to the energy loss of plasma in the two-temperature region of the jet. We find that the luminosity of the high energy photons emitted from the region outside the sonic point amounts to ∼ 2× 1039 ergs s−1, which is approximately comparable to the kinetic luminosity of the outflowing matter in the jet beam. In the n = 1 case electrons lose energy mainly by adiabatically expanding. In Table 3 the adiabatic cooling is adopted for the n = 1 case. The sonic point is located close to the central engine. As in the n = 2 case the analytic solution also yields the electron temperature an order of magnitude lower than the ion temperature and large luminosity of high energy photons. Note, however, that at the sonic point the radius of the jet cross section Rs is an order of magnitude larger than the distance rs. Although assumed in the calculation, the uniformity of physical parameters over the jet cross section may not be guaranteed when R ≫ r. Hence, the physical parameters at the sonic point, shown in Table 3, may not be reliable. We expect that the analytic solution for the n = 1 case can be applicable down to the point where R ∼ r. We list the physical quantities there in the Table 4. The two-temperature solution for the n = 1 case can extend down to the distance r ∼ 2.3 × 109 cm and cannot reach the sonic point keeping R < ∼ r. Let us consider the hybrid jet model combining the n = 1 and n =2 cases. We adopt the n = 1 case presented above to describe the outer part of the jet beyond the distance r ∼ 2.3 × 109 cm, while for the inner part we adopt the n = 2 case (conical jet). The conical jet extends inwards starting from the boundary with the physical quantities shown in Table 4. We approximate the electron cooling rate by the radiative loss rate. The density and ion and electron temperatures increase with decreasing distance, following equations (4), (24) and (18). We reach the sonic 10 [Vol. , point at the distance rs ∼ 9.2 × 107 cm and derive physical quantities such as shown in Table 5. Note that the sonic point in the hybrid case locates much closer to the central engine compared to the simple conical case shown Author(s) in Table 3. The luminosiy of high energy photons emitted from the region outside the sonic point is also very large and amounts to an important fraction of the kinetic luminosity of the outflowing matter. The other properties of the two-temperature jet are more or less similar to those of the simple conical case. Fukue (1987) already considered the hybrid model for the SS 433 jets. Contrary to our hybrid model, he adopted the n = 1 and n = 2 beams respectively for the inner and outer parts of jets. The sonic points lie close to the compact objects similarly to our calculation for the n = 1 case (Table 3). The boundary between the n = 1 and n = 2 beams may roughly correspond to the base of the X-ray emitting region. If we adopt a standard set of parameters for the physical quantities at the base (Kotani et al 1996), we obtain the radius of the jet cross section at the sonic point which exceeds the distance of the sonic point from the compact stars, especially in the neutron star case. The uniformity of the physical parameters, mentioned above, may not be guaranteed also in the hybrid model by Fukue (1987). We note that the electron-positron pairs and appropriate compton amplification in addition to the pressure gradient force should be included in the calculation in order to have more accurate estimates on the jet property especially around the sonic point. Even if included, the qualitative properties may not deviate significantly from those obtained here. 5. Discussion and Conclusions The jet beams in SS 433 may originate near the compact object, or more specifically in the inner region of the accretion disk. The X-ray emitting region of the jet is likely to be located at the distance of > ∼ 1011 - 1012 cm from the compact object. The jet beams must be accelerated and collimated to reach the terminal velocity within the distance of < ∼ 1011 - 1012 cm. The mechanisms of acceleration and collimation for the jet beams in SS 433 are poorly understood. In general the gas and radiation pressures and magnetic field are considered as a possible driving force for the jet acceleration. If the shape of the SS 433 jets is simply conical in the X-ray emitting region and also in the inner region adjacent to it, we find that the sonic point should be located very far from the central and compact object. The ion temperature at the sonic point is very high and close to the maximum attained near the compact object. It is quite difficult to No. ] Running Head 11 keep the jet matter so hot when the matter outflows from the jet footpoint near the central object to the distant sonic point. The jet matter will suffer from the adiabatic cooling and lower the tempetature substantially unless the shape of the jet is completely columnar in the region from the footpoint to the sonic point. Let us turn to the jet acceletation in this conical case. The free fall velocity at the sonic point due to the gravity of the compact object is more than an order of magnitude smaller than the jet velocity. Hence, the acceleration due to the gas pressure may not be applicable in this case. The radiative acceleration is also ruled out since the inner and hotter portion of the jets is optically thick along the flow direction and hence photons from outside cannot enter into the jet and impart momentun to the matter. The magnetic driving still remains as a possible acceleration mechanism. These arguments indicate that the simple conical model for the SS 433 jets is less likely. Promising is the hybrid model presented in the previous section. We adopt the model jet with n = 1, where the cross section of the jet grows with the distance more slowly than that of the conical jet, in order to describe the outer part of the jet including the X-ray emitting region. We use the conical jet to describe the inner and hotter part of the jet. This model explains the properties of observed X-rays well. Furthermore, in the hybrid model the sonic point can come close to the central object. Hence, the gas pressure as well as the magnetic force may also be considered as a possible driving force for the jet acceleration. The radiative acceleration is again ruled out since the inner and hotter portion of the jets is optically thick along the flow direction. We note that the inner and hotter portion of the jets emit a huge amount of gamma rays. No observational evidence, however, has not been found for an emission of such high energy photons. Hence, we conclude that the inner and hotter portions of the jets are hidden by the surrounding matter, most probably by the thick envelope and accretion disk. Note that such an envelope, or a dense outflowing atmosphere which envelops the system, is also suggested from the substantial fluctuations of the optical magnitudes and the infrared spectrum (Vermeulen 1989; Band 1987). The collimation of the jet beams is another important issue in the study of the astrophysical jet. The hybrid model above indicates that the jet beam may be squeezed and collimated more strongly around the distance of ∼ 109 cm. We expect that at squeezing a part of the kinetic energy of the jet matter may be dissipated into heat and reheat the jet matter. A substantial amount of surrounding matter is also required, again, if it is responsible for squeezing, from the pressure balance with the ram pressure of the jet matter. If the thick accretion disk channels the jet beam and precesses, the precession of the jet would be a natural consequence. The magnetic confinement is another possibility for collimation. Our studies show that the magnetic fields, which yield the magnetic pressure 12 Author(s) [Vol. , comparable to the gas pressure at the sonic point, are of the order of 107 G. This field strength may also be used to study and set constraints on the properties of the magnetically driven jets. We thank N. Kawai and T. Kotani for useful discussions. We also thank the anonymous referee for useful comments. This research was supported in part by the Grant-in Aid for Scientific Research (C) (10640234, 12640302). References Band D. L. 1987, PASP 99, 1269 Kotani T., Kawai N., Matsuoka M., Brinkmann W. 1996, PASJ 48, 619 Kotani et al. 1997, in Proceedings of the Fourth Compton Symposium, Ed C. D. Dermer, M. S. Strickman, J. D. Kurfess, AIP Conference Proceedings 410, 922 Fukue J. 1987, PASJ 39, 679 Margon B. 1984, ARA&A 22, 507 Vermeulen R. C. 1989, Ph. D Thesis, University of Leiden No. ] Running Head 13 Table 1. Physical quantities at the base of the X-ray emitting region. n r0 R0 ρ0 T0 ξ0 M L(x0) Lk (1010cm) (1010cm) (10−11g cm−3) (108K) (10−6M⊙y−1) (1036ergs s−1) (1039ergs s−1) 1 2 6.4 52 1.2 2.3 4.1 1.5 2.3 0.05 2.3 0.15 2.3 3.1 1.0 0.97 4.4 5.9 Table 2. Physical quantities at the boundary inside of which the jet plasma is in the two-temperature regime. n r1 R1 ρ1 T1 ξ1 L(x1) (1010cm) (109cm) (10−10g cm−3) (108K) (1036ergs s−1) 1 2 1.3 12 5.5 5.5 2.0 2.6 6.9 0.67 20 1.6 1.9 12 Table 3. Physical quantities at the sonic point. n rs Rs ρs Ts Tes ξs L(xs) (106cm) (108cm) (10−8g cm−3) (1011K) (109K) (1039ergs s−1) 1 2 4.3 1.0 1.1 × 104 5.2 61 2.9 4.4 9.1 0.014 0.084 4.4 17 0.64 3.8 Table 4. Physical quantities at the distance where R = r in the n = 1 case. r R ρ T Te ξ L(x) (109cm) (109cm) (10−9g cm−3) (109K) (109K) (1037ergs s−1) 2.3 2.3 1.1 4.2 1.4 0.23 1.6 14 Author(s) [Vol. , Table 5. Physical quantities at the sonic point in the hybrid model. rs Rs ρs Ts Tes ξs L(xs) (107cm) (107cm) (10−7g cm−3) (1011K) (1010K) (1039ergs s−1) 9.2 9.2 6.9 4.4 1.7 0.13 1.3
astro-ph/0004077
1
0004
2000-04-06T09:40:15
Interpretation of the UV spectrum of some stars with little reddening
[ "astro-ph" ]
The spectrum of five stars is analysed and explained by the superimposition of two components. One is the extinction of the direct starlight, with an extinction coefficient A which varies as the inverse of the wavelength across all the UV. This linear dependence of extinction upon the inverse of wavelength prolongs to the UV the well known relation which exist in the optical between extinction and wavelength. The second component is an additional feature, superimposed on the extincted direct starlight. This feature is interpreted as starlight scattered by dust at close angular distance to the star, into the beam of the observation.
astro-ph
astro-ph
Interpretation of the UV spectrum of some stars with little reddening. Fr´ed´eric Zagury Department of Astrophysics, Nagoya University, Nagoya, 464-01 Japan email: [email protected] December, 1999 One is the direct starlight, F 0 Abstract. The UV spectrum of a few reddened stars will be decomposed into two terms. ⋆,λe−τλ, which is the product of the flux of the star corrected ⋆,λ, and of the extinction e−τλ. The second is starlight scat- tered by interstellar dust into the beam of the observation. This excess of scattered for interstellar extinction, F 0 starlight affects the FUV part of the spectrum (λ < 2200 A). The combination of both terms gives the shape of the UV spectrum of a reddened star, with its characteristic depression at 2200 A. 1. Introduction This paper is the second of a serie dedicated to the interaction of starlight and interstellar grains in the UV and supported by the observations of the International Ultraviolet Explorer (IUE) satellite. The properties of the interstellar grains can be studied either by their capacity to scatter starlight, by observations of reflection nebulae, or by their capacity to extinguish starlight, by observing a star through an interstellar cloud. The former method allows the evaluation of the phase function and of the albedo of interstellar grains. The latter gives more specific information on the proportion of starlight to be absorbed or scattered at each wavelength. If extinction is the only process involved, the curve which is obtained by dividing the spectrum of a star by the spectrum of an unreddened star of same spectral type shows the relative capacity, from one wavelength to another, of the grains to extinguish starlight in the direction of the star. This is an intrinsic property of the interstellar grains present in the direction of the observation. The most salient feature of the UV spectrum of a star is the 2200 A bump which appears when there is interstellar matter between the star and the observer. From the Send offprint requests to: F. Zagury 2 Fr´ed´eric Zagury: Extinction in the UV-II standpoint of interstellar grain properties, and if no scattered starlight is introduced into the beam of the observation, this implies the existence of a particular class of grains, the bump carriers, which extinguish light at wavelengths close to λb ∼ 2175 A. Those particles have never been formally identified to date. Efforts to comprehend the variations of the stars UV spectrum in different interstellar environments, all of which have assumed starlight extinction as the only process involved, did not really succeed in bringing a global understanding of the UV extinction curve (see Savage et al. 1985, Fitzpatrick & Massa 1986 and 1988). In a preceding paper (Zagury 2000, paper I hereafter) the existence of the bump carriers was questioned as these carriers do not affect the UV spectrum of reflection nebulae. The UV spectrums of the bright nebulae presented in paper I were all interpreted as the result of starlight scattered by interstellar grains with identical albedo and phase function across the UV and in the different directions of space sampled by the nebulae. It was also noted that these properties of the grains may be identical in the optical spectral range. No bump or other particular feature is created at 2200 A in the spectrum of the light scattered by a nebula. If the bump carriers are not present in nebulae, the presence of an additional compo- nent due to scattering becomes the most reasonable explanation of the bump (Bless & Savage 1972, and the appendix of this paper). To explain the variations of the surface brightness of a nebula in function of its distance θ from the illuminating star (paper I), the interstellar grains must have a strong forward scattering phase function. In paper I it was found that the maximum surface brightness a nebula can reach varies as a power law θα (α < −1) of θ. A value α = −2 was found. Consequently, the starlight scattered into the beam of observation at close angular distance to the star may be a non-negligible propor- tion of the direct starlight. The scattered starlight will be more important in a wavelength range for which the scattering medium has an optical depth close to τmax ∼ 1 − 2. The present paper will further develop these ideas. I will study the UV spectrum of a selected sample of stars with a bump and little reddening. Contrary to previous studies which use the logarithm of the spectrums as a mean of obtaining the extinction curve, in this paper, the direct linear data will be used. This is necessary if the spectrums can be decomposed into two separate components, the direct starlight and the scattered starlight. The stars to be studied in this paper have been selected because of the simple and straightforward interpretation which can be given of their UV spectrum. The specific aspect of the UV reduced spectrum of these stars, defined as the ratio of the star to an unreddened star of same spectral type spectrums (section 2), clearly separates two spec- tral regions. The long wavelength part of the spectrum is correlated with the reddening of Fr´ed´eric Zagury: Extinction in the UV-II 3 the star, E(B − V ), and, for some stars, it is fitted down to the bump spectral region by an exponential of 1/λ. The exponential will be interpreted as the extinction of starlight e−τλ, where τλ is a linear function of 1/λ (section 4). The linear in 1/λ dependence of τλ was established in the optical in the 1930's (Greenstein & Henyey 1941 and references therein) and detailed in more recent studies by Rieke & Lebofsky (1985) and Cardelli, Clayton and Mathis (1989). For the stars I have selected, this law extends to the UV. At shorter wavelengths, λ < λb, a bump-like feature appears, superimposed to the exponential decrease. This feature will be analysed in section 6 and attributed to ad- ditional starlight scattered by interstellar dust at very small (compared to the beam of the observations) angle to the star. Hence, the 2200 A bump is no longer considered as a depression, but rather the point at which scattering becomes noticeable. The consequences of this interpretation are discussed in the conclusion. 2. Data The IUE experiment (Boggess et al. 1978a, 1978b) and the process employed to obtain the final IUE spectrums have been described in paper I. This paper will be concerned with seven stars, listed in table 1. Two of these stars, HD23480 (Merope) and HD200775 illuminate the Merope nebulae and NGC7023, which have been studied in paper I. For each star, the spectrum presented in the paper is an average of the best observations of the object. Some stars having been observed many times, only a few observations were necessary to ensure sufficient accuracy. At times I used high dispersion spectrums and decreased the resolution by a median filter. The spectrum of the seven stars selected for this study are presented in figure 7, as they can be seen at IUE website. The ratio of a reddened star spectrum to the spectrum of an unreddened star of same spectral type (reference star) will be called a 'reduced spectrum' of the reddened star. Each star has many reduced spectrums. All reduced spectrums of a star are proportional. The reduced spectrum of the seven stars, scaled to a common value at 3 µm−1, are presented in figure 2. The 'absolute reduced spectrum' of a star is the reduced spectrum of the star ob- tained if the reference star is the star itself, corrected for reddening. It characterizes the interstellar medium between the star and the observer (section 3). The logarithm of the absolute spectrum of a star is proportional to the extinction curve, Aλ versus 1/λ, in the direction of the star. The reference stars which have been used to establish the reduced spectrum of the stars are listed in table 2. 4 Fr´ed´eric Zagury: Extinction in the UV-II Fig. 1. Left : Spectrums of unreddened stars with close spectral type. Right : Variations of the UV spectrum of unreddened stars according to the spectral type. A star spectrum can be presented as a function of the wavelength λ or as a function of λ−1. The former manner keeps close to the data, whereas the latter is preferable since the optical depth varies as 1/λ. The latter presentation, more useful when dealing with scattering, has been chosen for most plots. 3. The 2200 A bump and the extinction of starlight When only extinction affects the transport of light between a star and the observer the spectrum of a star is the product of two terms: the unreddened flux of the star F 0 ⋆,λ and the extinction e−τλ, where τλ = 0.92Aλ is the optical depth at wavelength λ of the medium in front of the star. A reduced spectrum of the star, proportional to e−τλ, depends only on the interstellar medium between the star and the observer. The absolute reduced spectrum of the star is e−τλ. The presence of the bump has complicated the interpretation of the UV spectrum of the stars but has not changed the ideas presented so far: the bump is usually considered as a particular feature in the τλ function. The bump was proven to originate in the interstellar medium and related to the quantity of interstellar matter in front of the star (Savage 1975, see also Savage, Massa and Meade 1985). Stars of same spectral type with a bump have significant differences in their U.V. spectrums while stars with no bump and close spectral types superimpose very well after multiplication by an appropriate factor (figure 1). 4. The exponential decrease Figure 2 plots the reduced spectrums of the seven selected stars. The reduced spectrums are smoothed by a median filter and scaled to have the same value at 3 µm−1. Stars are listed on the figure by order of increasing E(B − V ), written after the star name. Fr´ed´eric Zagury: Extinction in the UV-II 5 Fig. 2. Reduced spectrums of the stars divided by their value at λ−1 = 3 µm−1. Fig. 3. left plot : Logarithm of the reduced spectrum of the stars of figure 2 in the long wavelength range. right plot : Correlation between E(B − V ) and the exponent of the exponential decrease. The solid line has a slope of 2. Each spectrum consists of 2 parts, clearly separated at λb. The long wavelength part (3 µm < 1/λ < 1/λb) varies steeply with increasing 1/λ. This decrease will be called the 'exponential decrease'. 6 Fr´ed´eric Zagury: Extinction in the UV-II There is a close correlation between the E(B − V ) value of a star and the steepness of its' exponential decrease: stars with larger E(B − V ) have a more rapid decrease. For all the stars of the sample, the exponential decrease is well fitted by an expo- nential, Ex = βe−e/λ, of 1/λ. The exponent e can be estimated by taking the logarithm of the reduced spectrums (figure 3, left). In some directions (HD147889, BD +62 2154), especially the directions of highest E(B − V ), the exponential fit is best at long wave- lengths, close to the optical wavelengths. In other directions, e.g. the directions of lowest E(B − V ) (HD23480, HD37903, HD149757, HD62542, HD200775), the fit applies to the totality of the exponential decrease. Figure 3, right, plots e as a function of E(B −V ). Within the error margin of E(B −V ) (estimated to be ∼ 0.1 mag) and on the determination of e (∼ 5%), we have: e ∼ 2E(B−V ). This relation is justified if the linear relation which holds between Aλ = 1.08τλ and 1/λ in the visible (Cardelli, Clayton & Mathis 1989, Rieke & Lebofsky 1985) is extended to the UV. In this case: Aλ = E(B − V ) 1 λB − 1 λV ( 1 λ = 2.2E(B − V )( τλ = 2E(B − V )( + 0.46(RV − 4)) ) + AV + 0.46(RV − 4)) − 1 λV 1 µm λ 1 µm λ (1) (2) with R−1 V = (AB − AV )/AV = τB/τV − 1. In the spectral range where equation 2 applies, extinction decreases as e−τλ ∝ e−2E(B−V )/λ. The Ex functions which fit the near UV spectrum of the stars we are concerned with have an identical exponential dependence on 1/λ (exponent = −2E(B − V )/λ). The Ex function prolong the linear optical extinction in the UV. Since there is no reason to suspect an abrupt change in the optical depth -more specifically in the constant term of equation 2- when going from the optical to the UV, relation 2 must hold, in the directions sampled by the stars, from the near infrared to the near UV (λ > λb). The left plots of figures 4, 5 and 6-top represent a reduced spectrum in the directions where the exponential decrease can be fitted by the Ex function down to λb. The solid line is the exponential fit. The close fit the Ex function provides to the exponential decrease indicates that, in these directions, the extinction curve is a linear function of 1/λ according to equation 1 from the optical to λb. 5. Absolute reduced spectrums Knowledge of E(B − V ) along with relation 2 permits an exact scaling of the reduced spectrums in the directions where the exponential decrease is well fitted by an exponential function of 1/λ. Fr´ed´eric Zagury: Extinction in the UV-II 7 Fig. 4. Figure 4 to figure 6. left plot : reduced spectrum of the star and best exponential fit to the long wavelength part of the spectrum. The curve in dots is the unsmoothed spectrum. The solid line spectrum was obtained by applying a median filter to the un- smoothed spectrum. The right y-axis is the absolute calibration of the plot, assuming RV = 3. right plot : Absolute reduced spectrum of the star minus the exponential de- crease. 8 Fr´ed´eric Zagury: Extinction in the UV-II Fig. 5. Same as in figure 4. Fr´ed´eric Zagury: Extinction in the UV-II 9 In these directions, the exponential Ex = βe−2E(B−V )/λ multiplied by a scaling factor αs must equal the extinction of starlight, e−τλ. αs is unambiguously determined from equation 2 by: αs = = 1 β 1 β e−0.92E(B−V )[RV −4] e−0.92AV (1− 4 RV ) (3) (4) αs is the factor to be applied to the star reduced spectrum in order to obtain the absolute reduced spectrum and the extinction curve in the direction of the star. In low density regions where RV is supposed to be ∼ 3 (Cardelli et al. 1989), we have: τλ = 2E(B − V )( 1 µm λ − 0.46) αs = 1 β e0.92E(B−V ) (5) (6) With equation 6 the absolute reduced spectrum in the direction of a star can be deduced from a reduced spectrum and from the associated Ex function. This property was used to calibrate the right axis of the left hand plots and the right hand plots of figures 4 to 6. 6. The FUV spectrum 6.1. Analysis There is no evident correlation between E(B − V ) in the direction of a star and the short wavelength part (1/λb < 1/λ < 8 µm) of its reduced spectrum (figure 2). The bump-like feature at 1/λ > 1/λb is added to the tail of the exponential decrease. When the latter is slow, indicating small E(B − V ), the bump feature is tilted because of the underneath exponential decrease. HD23480, HD149757, HD62542, HD200775 are typical examples (see figure 2 and the left plots of figures 4 and 5). In the directions of the stars selected in figure 4 to figure 6, the Ex function fits the exponential decrease down to λb, showing no excess of extinction or other peculiarity at this wavelength. In all directions the reduced spectrum seems to catch up with the exponential decrease for large τλ-values. The reduced spectrum of the stars, right hand plots of figure 4 to figure 6, comprises the expected exponential extinction of starlight and the additional bump-like feature at short wavelengths. The bump feature can be isolated if, for each direction, e−τλ is substracted from the absolute reduced spectrum in the same direction. The resulting curves are plotted at the right hand of figure 4 to 6. The short wavelength bump in the different directions are represented on the same plot, figure 6, down and left. With increasing E(B − V ), the height of the bumps tends to decrease and the short wavelength decrease (high τλ) is steeper. 10 Fr´ed´eric Zagury: Extinction in the UV-II 6.2. Scattering Scattering was ruled out as an explanation of the 2200 A bump for reasons which are summarized in the appendix and are questionable. The spectrum of nebulae (paper I) did not reveal the presence of the bump carriers in the interstellar medium. The expo- nential decrease of the UV spectrums can also logically be interpreted as the extinction of starlight by the same mean grain population responsible of the 'normal' extinction process of starlight. For the stars studied here there does not seem to be any peculiarity at λb, and there is no need of a particular class of grains which would extinguish starlight at this wavelength. The bump-like FUV feature is superimposed on the exponential de- crease and appears as an additive feature to the extinction of the direct starlight e−τλ. It must arise (Bless et Savage 1972 and the Appendix, this paper) as a result of the introduction of scattered starlight into the beam of observation. a star is the sum of the direct starlight, F 0 ⋆,λScaλe−τ ′ F 0 (relative F 0 If scattering is added to direct starlight, the light we receive from the direction of ⋆,λe−τλ, and of the scattering component ⋆,λ is the stellar flux at λ corrected for reddening, Scaλ is the proportion λ < τλ is the optical depth which accounts for the extinction of starlight between the star and the scattering ⋆,λ) of starlight at wavelength λ scattered into the beam. τ ′ λ. F 0 medium and for the extinction of the scattered light. The Scaλ function depends on the structure of the medium sampled by the line of sight. The light scattered by a medium made of clumps with low density will not have the same spectrum as the light scattered by high density clumps. The small scale structure of the medium, that is the density distribution of the different regions which compose the scattering medium at scales probably smaller than the IUE beam (Falgarone et al 1998), determines the shape of the Scaλ function. With increasing E(B − V ) the FUV bump has a smaller maximum and a steeper FUV decrease (left-bottom plot of figure 6). This can be attributed to the e−τ ′ λ term of the scattered emission which affects the shortest wavelengths. In the right hand bottom plot of figure 6 the FUV bumps are multiplied by an appropriate function e−τ ′ λ given by equation 5 with E(B − V ) = 1, and γ an appropriate constant. γ is adjusted for all the curves to have comparable maximums. λ = e−γτ 0 λ, τ 0 All the curves of the figure have an identical growth between λb and the maximum. The curves for the direction of lowest reddening (HD23480, HD37903, HD149757) superimpose well, supporting the idea of scattering by a medium with similar character- istics in these directions: the Scaλ functions are identical and the scattered component of the absolute reduced spectrums differ by the extinction term e−τ ′ λ only. Fr´ed´eric Zagury: Extinction in the UV-II 11 The curves for the directions of HD200775 and HD62542 cannot be superimposed to the others, thus differing by the Scaλ function, which indicates an environment of different nature. 6.3. Implications If scattering is responsible for the FUV bump of a star, the bottom plots of figure 6 show that it can represent up to ∼ 15% of the direct starlight received on earth and corrected for extinction. Most of the additional light has to be scattered within the θ0 = 1" angle of the small S aperture of the IUE telescope since the differences between observations made with the S and with the large L apertures of the IUE telescop are explained by pointing problems (section A.1.). Because the ratio of L to S observations is generally greater than 1.2, the difference of the amount of scattered starlight between the two types of observations must be less than 20%. Thus, if scattering is responsible for the FUV bump of a star, the scattered light is necessarily an appreciable proportion of direct starlight and scattering by interstellar dust must be strongly oriented in the forward direction. 6.4. Case of a θ−2 dependence of the maximum surface brightness of nebulae Suppose the θ−2 dependence of the maximum surface brightness of a nebula (paper I) on angular distance θ to the illuminating star is verified and applies to very small angles, down to a lower limit θmin. According to the equation 6 of paper I, the amplitude of the FUV bump, 15% of the unreddened flux of the star measured on earth, will be justified if: θmin < θ0e− 0.07 πc θmin < 6 10−4 " (7) The latter inequality assumes c ∼ 3 10−3 and θ0 = 1 ". Note that this upper estimate of θmin is extremely sensitive to the c parameter. A value of c = 15 10−3, within the possible range of values observed in paper I, gives θmin < 0.2 ". c = 10−3 gives θmin < 2 10−10 ". If ra is the ratio of the maximum light received from the direction of a star observed with apertures θ1 and θ0: ra = ∼ πc(1 + 2 ln(θ1/θmin)) πc(1 + 2 ln(θ0/θmin)) ln(θ1/θmin) ln(θ0/θmin) ∼ 1 − ln θ1 ln θmin (8) 12 Fr´ed´eric Zagury: Extinction in the UV-II with θ1 and θmin in arcsecond. If θ1 ∼ 10", corresponding to the large aperture of the IUE telescope, and θmin is of order 10−α ", equation 8 has the simple form: ra = 1 + 1 α (9) Within this framework, conditions 7 and 8 refine the validity domain of the θ−2 law. Condition 7 must be satisfied to account for the amount of scattered light which is observed. The power received in the direction of a star will not differ from the small (θ0 ∼ 1 ") to the large (θ1 ∼ 10 ") aperture of the IUE telescope if α ≫ 1 (equation 9). ra less than 1.2 will be achieved for α ≥ 5, i.e. θmin ≤ 10−5 ". In general, for observations made with an aperture θ1 ∼ 10αm ", and provided that the θ−2 dependence of the surface brightness extends to 10αm ", ra will be 1 + αm/α. This result also depends upon the filling factor of the scattering medium. 7. Conclusion The UV spectrum of selected stars with a bump and moderate reddening was decomposed into two parts. The first part is the expected extinction of starlight, e−τλ, with τλ a linear function of 1/λ. The second is the FUV bump, observed in figure 2, which is superimposed on the tail of the exponential decrease. This additional component was interpreted as starlight scattered by interstellar dust at very small angle to the star. This decomposition is justified by the very close exponential fit which can be applied to the long wavelength exponential decrease of the spectrums (figure 4 to 6). The ex- ponent of the exponential is e = 2E(B − V )/λ, as in the optical. The exponential fit extends to the UV the linear relation between Aλ and 1/λ which applies in the optical. There is no excess of extinction at 2200 A in the spectrum of the stars which have been selected. The additive nature of the FUV bump follows from the analysis of figure 2 carried out in sections 4 and 6. Its' interpretation as scattered starlight is then the most credible one. This interpretation provides a simple explanation of the similarities and of the differences between the bumps in the different directions (section 6.2). If scattering is present in the spectrum of the stars, most of it must occur at less than 1 " to the star. It implies a strong forward scattering phase function of the interstellar grains, as found in paper I. An attempt to justify the amount of scattered light (∼ 15% of the star flux measured on earth and corrected for reddening) was proposed in section 6.4. It involves the power law found in paper I, which need to be confirmed, of the variation of a nebula maximum surface brightness with angular distance to the illuminating star. This interpretation of the UV spectrum of the stars has two important consequences. Fr´ed´eric Zagury: Extinction in the UV-II 13 If the UV spectrum of some stars with a bump is explained without requiring an excess of extinction from the bump carriers at λb, all the stars' UV spectrum must have a similar interpretation. There is no reason why scattering should affect the spectrum of some stars solely and/or why the bump carriers should be present in some nebulae only. In the directions of the selected stars, the extinction curve, Aλ as a function of 1/λ, is a straight line from the near infrared to the FUV. The value of Aλ at wavelength λ is given by the equation 1. If, as it was suggested in paper I, grains have the same properties in all directions of space, this law must hold for all directions. Why is scattering so important in the FUV? A first reason comes from the extinction of starlight which is increased when moving to the shortest wavelengths. Consequently, at short wavelengths, the scattered starlight will be a larger part of the total light received in the direction of a star. It is also plausible that the structure of the interstellar medium favors scattering at particular wavelengths. A medium constituted of small clumps of similar AV will scatter starlight in a particular wavelength range: high AV clumps will preferentially scatter toward the red while very low column density clumps will scatter in the UV. Although this aspect of scattering was not developped here, it may be an important step in the comprehension of the spectrum of the stars. 14 Fr´ed´eric Zagury: Extinction in the UV-II Fig. 6. upper plots: Same as in figure 4. bottom left : The absolute reduce spectrum of the stars are presented on the same plot. bottom right : The absolute reduce spectrums are scaled to a similar maximum value (∼ 0.15). The spectrum are multiplied by an e−τ ′ λ function. τ ′ λ was empirically determined using equation 2 and RV = 3. Fr´ed´eric Zagury: Extinction in the UV-II 15 Fig. 7. UV spectrum of the reddened stars used in the paper. 16 Fr´ed´eric Zagury: Extinction in the UV-II name α1950 δ1950 l b B B − V E(B − V ) sp. type P ar a1 ⋆ a2 st BD +62 2154 22 58 33 +63 14 52 110.94 +03.26 HD147889 16 22 23 −24 21 07 352.86 +17.04 9.76 8.66 0.43 0.71 HD149757 ζ Oph 16 34 24 −10 28 02 62.8 +23.59 2.595 0.017 HD200775 21 01 00 +67 57 55 104.06 +14.19 7.73 0.31 HD23480 Merope 03 43 21 +23 47 39 166.57 -23.75 4.113 -0.051 HD37903 HD62542 05 39 07 −02 16 58 206.85 -16.54 7 40 58 −42 06 37 255.92 -09.24 7.91 8.21 0.07 0.18 0.68 0.96 0.33 0.55 0.09 0.32 0.38 B1V B2III/IV O9V B2Ve B6IVe B1.5V B3V 7.36 7.12 2.33 9.08 2.12 4.06 6 3 1 5 2 6 4 a1 parallaxe in marcsec measured by Hipparcos a2 associated standard star number. Refers to table 2 Table 1. Stars used in the paper. Except for notified exceptions, all informations come from Simbad database (http://simbad.u-strasbg.fr). (B − V )0 from FitzGerald 1970 N◦ name α1950 δ1950 l b B B − V E(B − V ) sp. type P ara 1 2 3 4 5 6 HD214680 22 37 00 +38 47 22 96.65 -16.98 4.673 -0.2 0.1 HD215573 22 45 48 −80 23 19 309.03 -35.53 5.190 -0.123 0.017 HD31726 04 55 27 −14 18 28 213.5 -31.51 5.928 -0.206 -0.034 HD32630 05 03 00 +41 10 08 165.35 +00.27 3.012 -0.146 HD58050 07 21 37 +15 36 58 202.53 +14.19 6.261 -0.187 HD74273 08 39 30 −48 44 36 267.13 -04.27 5.69 -0.21 0.054 0.053 0.04 O9V B6IV B2V B3V B2Ve B1.5V 3 7.35 3.28 14.87 0.67 2.12 a parallaxe in marcsec measured by Hipparcos Table 2. Standard stars used for the stars in table 1. All informations come from Simbad database (http://simbad.u-strasbg.fr). (B − V )0 from FitzGerald 1970 Appendix A: The possible explanations of the 2200 A bump Bless & Savage (1972) review the possible causes of the bump: stars with a bump have a peculiar energy distribution; a large amount of starlight is scattered into the line of sight by interstellar dust; the extinction properties of the interstellar medium are particular in the UV. Because of its relation to E(B − V ), the bump does not originate in the stars' atmo- sphere, nor does it come from a special energy distribution. One explanation has been widely accepted and developped in all studies of the feature: the bump is due to a special class of interstellar grains which extinguish light at 1175 A. According to studies of the bump in various environments (Savage 1975, Jenniskens & Greenberg 1993, Nandy et al. 1976), those grains are well mixed (in all environments) to the large grains population responsible for the 'normal' extinction process of starlight. Fr´ed´eric Zagury: Extinction in the UV-II 17 The last possibility, that scattered light enters into the beam, was ruled out for reasons to be discussed in A.1.. A.1. Arguments used against scattering 3 types of arguments were used to rule out the possibility of scattering as an explanation for the bump. According to Bless & Savage (1972), Code has calculated that grains of albedo close to one are required to produce enough scattering to explain the UV extinction curve. However those calculations suppose isotropic scattering. The importance of forward scat- tering, and its' consequences for our interpretation of the UV spectrum of nebulae has been emphasized in paper I. Introduction of forward scattering will invalidate Code's calculations. Snow & York (1974) have compared the spectrums of σ Sco from 2 different observa- tions, each of which involves a camera of different aperture. The Wisconsin spectrometer aboard the Orbiting astronomical Observatory 2 (OAO-2) has a large aperture of 8′ × 3◦, while the Princeton OAO-3 has an entrance slit 0.3′′×39′′, 103 smaller than OAO-2. Thus, if the UV spectrum of stars with a bump was due to pollution by scattered nebular light, the authors expect to find differences between the 2 spectrums. No such difference is observed. Snow & York's work imposed a serious constraint on interstellar grains but did not demonstrate the abscence of scattering in the UV spectrum of stars. If the phase func- tion is strongly forward scattering, that is if the assymetry parameter g0 is close to 1, significant scattering will come only from directions very close to the star. Both observa- tions used by Snow & York have relatively large apertures, and will receive nearly equal amounts of scattered light. Snow & York's experiment can be repeated with IUE data. In most of the stars observed with both apertures of the IUE camera there is a scaling factor of 1.2 to 2 between the L and S aperture spectrums, regardless of the star reddening. This difference also affects unreddened stars and is probably due to the difficulty of holding the star within the beam when using the small aperture. Hence, if UV spectrums are affected by scattering, it must occur at angular distances less than 1.5" from the star. Witt & Lillie (1973) study the diffuse Galactic light (DGL) spectrum from the OAO-2 satellite. The arguments of this paper rely on models of both the interstellar medium, assumed to be a plane parallel slab of uniform density, and of interstellar grains. The apparent disagreement between the model and the observations is attributed to changes of dust albedo with wavelength and to a pronounced minimun of the albedo at 2200 A 18 Fr´ed´eric Zagury: Extinction in the UV-II (λb). No DGL spectra is presented in Witt & Lillie's paper but their results imply a pronounced minimum of the DGL at λb. IUE has observed over 400 off-positions, 'IUE SKY', free of luminous objects. Many of the observations have some cirrus on their line of sight, the 100 µm IRAS emission can range from 1 to a few 10 MJy/sr. The only region I have found with a reliable signal across the entire LWR camera wavelength range is spatially close to -and probably scatters the light of- the star cluster NGC1910. It has a level of 8 10−14 erg/cm2/s/A and no bump. In all other observations the emission shortward of 2600 A has a very broad amplitude which can be attributed either to noise or to a very low level of signal. None of the spectrums, even when many of them are co-added, shows evidence for extinction at 2200 A. References Bless R.C., Savage B.D, 1972, ApJ, 171, 293 Boggess A., et al., 1978, Nature, 275, 372 Boggess A., et al., 1978, Nature, 275, 377 Cardelli J.A, Clayton G.C, Mathis J.S, 1989, ApJ, 345, 245 Falgarone E., et al., 1998, A&A, 331, 669 FitzGerald M. P., 1970, A&A, 4, 234 Fitzpatrick E.L, Massa D., 1986, ApJ, 307, 286 Fitzpatrick E.L, Massa D., 1988, ApJ, 328, 734 Greenstein J.L., Henyey L.C., 1941, ApJ, 93, 327 Jenniskens P., Greenberg J.M, 1993, A&A, 274, 439 Nandy K., et al., 1976, A&A, 51, 63 Rieke G.H., Lebofsky M.J., 1985, ApJ, 288, 618 Savage B.D, 1975, ApJ, 199, 92 Savage B.D, Massa D., Meade M., 1985, ApJSup, 59, 597 Snow T.P., York D.G., 1974, Astrophys. Space Science, 34, 19 Whitford A.E, 1958, A.J., 63, 201 Witt A.N., Lillie C.F, 1973, A&A, 25, 397 Zagury F., 2000, submitted
astro-ph/9909474
1
9909
1999-09-28T20:26:48
NICMOS images of JVAS/CLASS gravitational lens systems
[ "astro-ph" ]
We present Hubble Space Telescope (HST) infrared images of four gravitational lens systems from the JVAS/CLASS gravitational lens survey and compare the new infrared HST pictures with previously published WFPC2 HST optical images and radio maps. Apart from the wealth of information that we get from the flux ratios and accurate positions and separations of the components of the lens systems that we can use as inputs for better constraints on the lens models we are able to discriminate between reddening and optical/radio microlensing as the possible cause of differences observed in the flux ratios of the components across the three wavelength bands. Substantial reddening has been known to be present in the lens system B1600+434 and has been further confirmed by the present infrared data. In the two systems B0712+472 and B1030+074 microlensing has been pinpointed down as the main cause of the flux ratio discrepancy both in the optical/infrared and in the radio, the radio possibly caused by the substructure revealed in the lensing galaxies. In B0218+357 however the results are still not conclusive. If we are actually seeing the two "true" components of the lens system then the flux ratio differences are attributed to a combination of microlensing and reddening or alternatively due to some variability in at least one of the images. Otherwise the second "true" component of B0218+357 maybe completely absorbed by a molecular cloud and the anomalous flux density ratios and large difference in separation between the optical/infrared and radio that we see can be explained by emission from either a foreground object or from part of the lensing galaxy.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 3 December 2017 (MN LATEX style file v1.4) NICMOS images of JVAS/CLASS gravitational lens systems N. Jackson, E. Xanthopoulos, and I.W.A. Browne NRAL Jodrell Bank, University of Manchester, Macclesfield, Cheshire SK11 9DL ABSTRACT We present Hubble Space Telescope (HST) infrared images of four gravitational lens systems from the JVAS/CLASS gravitational lens survey and compare the new in- frared HST pictures with previously published WFPC2 HST optical images and radio maps. Apart from the wealth of information that we get from the flux ratios and accu- rate positions and separations of the components of the lens systems that we can use as inputs for better constraints on the lens models we are able to discriminate between reddening and optical/radio microlensing as the possible cause of differences observed in the flux ratios of the components across the three wavelength bands. Substantial reddening has been known to be present in the lens system B1600+434 and has been further confirmed by the present infrared data. In the two systems B0712+472 and B1030+074 microlensing has been pinpointed down as the main cause of the flux ratio discrepancy both in the optical/infrared and in the radio, the radio possibly caused by the substructure revealed in the lensing galaxies. In B0218+357 however the results are still not conclusive. If we are actually seeing the two "true" components of the lens system then the flux ratio differences are attributed to a combination of microlensing and reddening or alternatively due to some variability in at least one of the images. Otherwise the second "true" component of B0218+357 maybe completely absorbed by a molecular cloud and the anomalous flux density ratios and large difference in separation between the optical/infrared and radio that we see can be explained by emission from either a foreground object or from part of the lensing galaxy. Key words: gravitational lensing - galaxies: active - galaxies: individual: B0218+357 - galaxies: individual: B0712+472 - galaxies: individual: B1030+074 - galaxies: indi- vidual: B1600+434 9 9 9 1 p e S 8 2 1 v 4 7 4 9 0 9 9 / h p - o r t s a : v i X r a c(cid:13) 0000 RAS 2 N. Jackson, E. Xanthopoulos, and I.W.A. Browne 1 INTRODUCTION Gravitational lens systems are important for a number of reasons. Individual lenses can be studied to determine a mass and information about the mass distribution of the lensing galaxy or cluster. Moreover, detailed studies of the lensing galaxies give clues to galaxy type and galaxy evolu- tion (Keeton, Kochanek & Falco 1998). If the lensed object is variable, measurement of time delays between variations of the multiple images, together with a good mass model of the lensing object, allow one to estimate the Hubble con- stant (Kundi´c et al. 1997; Schechter et al. 1997; Keeton & Kochanek 1997; Biggs et al. 1998). Statistical properties of lenses are also important. Both qo and λo can be constrained by lens statistics, a good example being the recent limit of λo < 0.66 for flat cosmologies (Kochanek 1996). For all these reasons, studies of larger samples of gravitational lenses are important. The Jodrell Bank VLA Astrometric Survey (JVAS) and Cosmic Lens All Sky Survey (CLASS) (Myers et al. 1998), in which 10,000 sources have so far been observed with the VLA, is the most systematic radio survey so far undertaken. It aims to look at all flat-spectrum radio sources in the northern sky with flux densities >30mJy at 5 GHz. Flat- spectrum radio sources are useful in this context for two main reasons; they have intrinsically simple radio structures, making the effects of lensing easy to recognise, and they are variable, allowing time delays to be derived. The ease of recognition means that this survey is statistically very clean and should not be subject to any significant selection biases -- important in view of the results to be outlined below. Here we present Hubble Space Telescope (HST) infrared imaging of four of the JVAS/CLASS lens sample. The main aim of this work is to obtain better constraints on the lensed images, in particular their flux ratios and their positions rel- ative to the lensing galaxy. Saha & Williams (1997) describe the importance of the use of the improved image positions and ratio of time delays between different pairs of images as a rigid constraint in the modelling of lenses especially when there is a limited number of observational constraints. They note however, that lately it has become apparent (Witt & Mao 1997) that in the case of quadruple-lens systems even well-determined image positions are not enough for accurate modelling, and thus they use non-parametric models. A second scope of the work is to compare the flux ra- tios of the components of the lens systems in the different wavebands (optical/near-infrared/radio). Variability of the source spectrum in connection with the time delay can lead to different flux ratios in different spectral bands. Another possibility is microlensing in at least one image. The stellar mass objects can magnify only sufficiently compact sources so if the flux ratios in two spectral bands in two or more images of a multiply-imaged QSO are different, this can be a sign of microlensing. For example, in the gravitational lens system 0957+561, one suspects microlensing effects from the different flux ratios in the optical continuum light, broad-line flux and in the compact radio component. A third possibil- ity is differential magnification, where parts of the source are more magnified than others. This can occur if the scale on which the magnification varies in the source plane is com- parable with the size of the emitting regions; so the effect can be different for radiation of different wavelengths. The most spectacular effects of microlensing occur when the an- gular radius of the source is much smaller than that of the Einstein ring for the microlens (Refsdal & Stabell 1997) but the effect has not been considered much up to now because of the long time scales that this occurs. However, the time scale is no more important when comparing the flux ratios of the different images in multiply lensed quasars. Mao & Schneider (1998) have suggested also that as in the case of microlensing for optical fluxes, substructure in the lensing galaxy can distort the radio flux ratios and this can be another cause of difference in the flux density ratios at different wavebands. Differential reddening can be an alternative explanation of differences observed in the flux density ratios of the lensed images at different wavebands. In gravitational lens systems, especially those where the images are seen through the lens, significant reddening by dust in the lens provides a natural explanation that must be considered first. There is consider- able evidence for dust reddening during passage through the lensing galaxy (Larkin et al. 1994, Lawrence, Cohen & Oke c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 NICMOS images of JVAS/CLASS gravitational lens systems 3 Object Obs date Time Observation number B0218+357 970817 B0712+472 970824 B1030+074 971122 B1600+434 970721 L S L S L S L S L S L S L S N44501KYM N44501KZM N44501L7M N44501L8M N44501LKM N44051LLM N44503GAQ N44503GBQ N44503GKQ N44503GLQ N44504VWQ N44504VXQ N44505DYM N44505DZM Table 1. Log of the observations. L indicates an exposure time of 2048s and S an exposure time of 576s. 1995; Jackson et al. 1998). Infrared pictures are less sub- ject to censorship by dust, generally giving the best view of the lensed images, and enable the degree of extinction to be quantified by comparison of brightnesses between images at different wavelengths. Another advantage of infrared obser- vations is that the contrast between host galaxy starlight and AGN continuum is less than in the optical. Thus, for not very luminous lens objects, we can expect to see arcs or rings formed by the extended host galaxy starlight. The lens system B1938+666 (King et al. 1998) is a particularly spectacular example of this phenomenon. Our aim is also to study the colours and light distributions of the lensing galaxies themselves. Keeton, Kochanek & Falco (1998) dis- cuss measurements of optical HST data on a number of the lensing galaxies, including B0218+357, B0712+472 and B1600+434, and use optical observations and lens modelling to address the question of dark matter distributions. Jackson et al. (1998) have published preliminary H-band photome- try of the JVAS/CLASS sample and show that all the lens systems discovered so far have lensing galaxies of normal mass-to-light ratios and that there is no evidence for dark galaxies in this sample. In section 2 we give details the HST observations. Sec- tion 3 contains brief descriptions and discussion of each in- dividual object, and section 4 contains the conclusions. 2 OBSERVATIONS All observations presented here were obtained with the Near Infrared Camera/Multi-Object Spectrometer (NICMOS) on the HST. Observations were taken with Camera 1, which has a pixel scale of 43 mas, and taken through the F160W filter, which has a response approximately that of the standard H- band. A list of observations, with exposure times and date of observations, is given in Table 1. Fits using the jmfit pro- gram in AIPS to the point spread functions generated by the TinyTim program (Krist 1997) yield full widths at half maximum (FWHM) of 131 mas for the NICMOS observa- tions. The complementary Wide Field & Planetary Camera 2 data presented in the Figures have FWHM of 65 mas and 79 mas for 555-nm and 814-nm images respectively. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The data were processed by the standard NICMOS cal- ibration pipeline at STScI. Multiple exposures were com- bined by weighted averaging or medianing depending on whether we had two or more images available. Bad pixels and cosmic rays were rejected by using a rejection algorithm. The images are first scaled by statistics of the image pixels, here the "exposure" for intensity scaling. The rejection al- gorithm used is the "crreject" which rejects pixels above the average and works even with two images. The algorithm is appropriate for rejecting cosmic ray events. As a second check we also added the images with binary image arithmetic within IRAF and divided the images by a number that is appropriate according to the exposures of the images and the number of the images involved so as to preserve the same counts/sec number as the initial NICMOS images. Cosmetically the result of both methods was the same. A point spread function (PSF) was generated by the TinyTim programme (Krist 1997) and deconvolved from the images using the allstar software, available within the NOAO IRAF package. In cases where the point sources lie on top of extended emission, such as B0712+472, the PSF deconvolution was done by hand, shifting and scaling un- til the smoothest residual map was obtained. Photometry and errors are based on the quantity of PSF subtracted and the range of PSF subtraction which gives a smooth residual map. In Fig. 1-4 we show the raw images and the images with the point-sources subtracted. 3 RESULTS AND DISCUSSION The general characteristics of all the four lenses can be found in Table 2. Columns 1 and 2 present the name of the survey in which each lens system was discovered and the name of the lens, while in columns 3, 4, 5, 6 and 7, one can find the number of components in the lens system, the image separation in arcsecs (the largest separation in the case of multiple components), the lens redshift, the source redshift and the morphology of the lensing galaxy respectively. 3.1 B0218+357 The B0218+357 system consists of two images of a compact radio source of a redshift 0.96 (Lawrence et al. 1995) sepa- rated by 334 mas. In addition to the compact images there is a radio Einstein ring (Patnaik et al. 1993). The lensing galaxy has a redshift of 0.6847 (Browne et al. 1993). It has been detected in V, I and H-bands with the HST (Fig. 1) and, as far as can be seen given the limits imposed by the signal in the images, it has a smooth brightness distribu- tion and is roughly circular in appearance. The colours of the galaxy are consistent with its previous classification as a spiral galaxy based on the presence of H1 absorption (Carilli, Rupen & Yanny 1993), strong molecular absorption (Wik- lind & Combes 1995) and large Faraday rotations for the images. It is worth noting that the presence of a rich in- terstellar medium means that extinction along lines of sight through the lensing galaxy may well be important, a possi- bility we will discuss below. There are several anomalies associated with the optical properties of the images in the B0218+357 system which 4 N. Jackson, E. Xanthopoulos, and I.W.A. Browne Survey Name # images ∆θ" zl zs lens galaxy JVAS B0218+357 CLASS B0712+472 JVAS B1030+074 CLASS B1600+434 2 + ring 4 2 2 0.334 1.27 1.56 1.39 0.6847 0.406 0.599 0.414 0.96 1.34 1.535 1.589 spiral spiral spiral spiral Table 2. The general characteristics of the four gravitational lens systems. Object Image Flux density Flux density Flux density (555 nm) (814 nm) (1.6 µm) 555 nm 10−20W m−2nm−1 Offset from A image 814 nm 1.6 µm (RA,δ,err) / mas Radio B0218+357 B0712+472 B1030+074 B1600+434 A B GAL A B C D GAL A B GAL A B GAL 2.0±0.5 13±1 6±2 1.02±0.10 0.33±0.03 0.40±0.04 <0.02* 4.9±0.5 2.0±0.5 19±2 13±2 0.87±0.09 0.33±0.03 0.38±0.04 0.02* 12±1 26.94±0.10 0.9±0.2 6? 27.37±0.07 1.17±0.17 7.9±0.7 3.9±0.3 2.0±0.2 * 2.9±0.3 2.2±0.2 9±2 12±1 19±1 15±2 0.98±0.15 0.41±0.06 0.41±0.06 0.22±0.05 11±1 47.6±0.4 1.40±0.09 10.8±0.5 0.75±0.1 0.53±0.1 4.5±1 - - - - 281,128,10 285,126,10 293,124,5 307,130 * - * - 66,−151,10 812,−659,5 64,−149,10 812,−660,5 * * 814,157,3 801,153,3 * - * 804,−642,10 1180,480,20 789,168,3 - - - - - 57,−160 812,−667 1163,460 - - 930,−1256,4 882,−1155,10 931,−1247,4 878,−1143,10 918,−1244,5 864,−1162,5 935,−1258 - - - - 726,−1188,3 726,−1184,3 708,−1193,10 723,−1189 * * * Table 3. Optical and F160W image photometry and positions. In each case the galaxy photometry refers to the light within the Einstein radius, apart from B0218+357 where the aperture was 3 square arcseconds. An asterisk indicates an image that is either invisible or impossible to deblend from another. Plate scales of 45.5 mas/pixel are assumed for the PC chip (555-nm and 814-nm images) and 43 mas/pixel for NICMOS. More detailed discussion of some of the optical images can be found in Jackson et al. (1998), Koopmans, de Bruyn and Jackson (1998) and Xanthopoulos et al. (1998) for B0712+472, B1600+434 and B1030+074 respectively. Errors in the radio positions are 1 mas or less. B1030+074A is approaching saturation on the NICMOS image. might be sufficient to doubt its classification as gravitational lens system, if it were not for the existence of an Einstein ring (Patnaik et al. 1993), a measured time delay (Corbett et al. 1995; Biggs et al. 1998) and high-resolution VLBA 15 GHz observations of B0218+357 (Patnaik et al. 1995). The observed anomalies are: • The optical and infrared flux ratios (see Table 3 and Grundahl & Hjorth 1995) for the images are very different from the well established radio ones (Patnaik et al. 1993). • The separation between the images may be less at opti- cal and infrared wavelengths than it is at radio wavelengths (see below). • The optical/infrared colours of the two images are very different. The V and I flux densities of B are consistent with the ground-based optical spectrum (Browne et al. 1993, Stickel et al. 1996; Lawrence et al. 1995). Similarly the H flux den- sity is what one might expect if B has a spectrum typical of a BL Lac-like object. The A/B flux density ratios are, however, unexpected. The most obvious result is that they are very different from the 3.7:1 measured in the radio (Pat- naik et al., 1993; Biggs et al., 1998). This could arise from micro-lensing of the optical/infrared emission and/or from extinction. However, extinction following a normal Galac- tic reddening law (Howarth 1983) is not consistent with the optical/infrared colours. This is because, if we attribute ev- erything to extinction, the necessary amount of reddening to give a 1:7 ratio in the V-band image should give about 1:2 in I, assuming an intrinsic 3.7:1 ratio of the radio im- ages. Even if we attribute the gross difference between the radio and optical/infrared ratios to micro-lensing, the fact that the V and I flux densities of image A are the same is not consistent with the A emission being a reddened version of the image B. Furthermore, if the flux densities for image B listed in Table 3 were severely contaminated with lensing galaxy emission, this still would not help. One last possibil- ity would be to invoke some variability in at least one image, although this would not explain the problems arising from comparison between the V and I images which were taken within an hour of each other. Hjorth (1997) has suggested that the separation be- tween the images may be less at optical and infrared wave- lengths (i.e. ∼300 mas) than it is at radio wavelengths (i.e. 334 mas). It is important to establish if this difference in image separations is real since, if true, it would immediately indicate that we are not seeing two images of the same object in the optical/infrared band. The radio image separation is well established to be 334±1 mas (Patnaik et al. 1993; Patnaik et al. 1995). The c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 NICMOS images of JVAS/CLASS gravitational lens systems 5 separations (and flux densities) given in Table 3 are the re- sult of by-eye fitting of point spread functions to the HST pictures in such a way as to obtain smoothest residuals. The optical pictures have much lower signal-to-noise than the in- frared ones, particularly for the A component and have cor- respondingly bigger errors. The image separation derived from fitting to the NICMOS picture is 318±5 mas, a value much smaller than the radio separation given above. This H- band separation is close to the one we find from the optical V and I WFPC2 images (308±10 mas and 311±10 mas in V and I respectively) but slightly larger than the optical sepa- ration found by Hjorth (1997) (296±10 mas and 299±10 mas from V and I respectively). The method we employ does not explicitly take account of the light from the lensing galaxy. Models consisting of an exponential disk plus point im- ages have been fitted to the data by McLeod et al. (1999). We have tried the same approach. A point spread function (PSF) was generated by the TinyTim programme (Krist 1997) and was fitted simultaneously to the A and B compo- nents of the lens system. We have tried fitting and subtract- ing the PSF for different positions of the two components. By attributing some of the emission around the B image (about 30%) to the core of the lensing galaxy (central 0.4 × 0.4 arcsec) we increase the separation until no flux is left (by subtracting both the components and the lensing galaxy) and so we find the maximum best fit to give a sepa- ration of 335 mas between the two components. The lensing galaxy has then an offset of 0.54,0.34 x, y pixels from the B component or it is 27 mas away at a position angle of 260◦ from B. By taking into account the light contribution from the lensing galaxy the method removes the discrepancy between the radio and optical/infrared separation. A limit can be deduced for the fraction of the light from the region of the B image which is contributed by the lensing galaxy. If we examine the optical spectra obtained at a number of different epochs (e.g. Browne et al. 1993, Stickel et al. 1996; Lawrence et al. 1995) they show no significant 4000A break nor any G-band absorption feature found in nearly all galaxy types (Bica & Alloin 1987). Of the light in the spectra ≤15% comes from the lensing galaxy. This implies that around the region of the B image, at an observed wavelength of ∼670 nm the galaxy contributes no more than ∼30% of the total flux density, allowing for slit losses and the fact that the A image makes a small contribution to the total light. Thus, if the lensing galaxy is as blue or bluer than the BL Lac-like spec- trum of lensed object, the model with a 30% contribution from the galaxy at 1.6 µm is only just consistent with the spectral data. Hence we conclude that, though a priori one would expect the radio and optical/infrared image separa- tions to be the same, there is some evidence to suggest that they are different. The anomalous flux density ratios may also in some way be related to the separation problem. So as mentioned above even if one takes into consid- eration the contribution of the light of the lensing galaxy to the B component there still is some doubt that the sep- aration problem between the optical/infrared and radio is completely resolved. Hjorth (1997) find from their optical data that the bright images that they observed could be identified with the radio A and B components because of the excellent agree- ment between the position angles, which is also in agreement with Grundahl & Hjorth (1995). They attribute the shorter c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 separation to the extendedness of component A and they argue that what we see is what is not covered by the molec- ular cloud (Grundahl & Hjorth 1995; Wiklind & Combes 1995). However a potential problem with this interpretation is that the bright image appears to be a point source and not extended as would be expected in this case. The alter- native explanation that they offer, that A is actually the core of the lensing galaxy while the counterpart to the B image is somehow swamped in the light of the galaxy core needs also to explain the fact that since B is a BLLac as proved from both ground based spectra and its V, I and H fluxes and the main dominant source of the optical contin- uum we are expecting to see a radio source at the location of the lens BL Lac (Urry & Padovani 1995) which we do not. Grundahl & Hjorth (1995) actually note that it is possible (due to the small intrinsic extent of A - 1 mas at high radio frequencies, so 5 pc at the redshift of the galaxy) that the molecular cloud can cover the entire image A. If this is the case then we are left with the difficult question concerning the nature of the A emission (hereinafter we refer to the op- tical/infrared emission near A as A∗) if it is not an image of the AGN. The fact that the true A image is obscured does not mean that all the light from host galaxy of the lensed object must be hidden too; some of its emission could be lensed and give rise to A∗. A difficulty with this idea is that the A∗ image is compact and of high surface brightness not extended as one might expect if it were an image of some part of the AGN host galaxy. Also there is no sign of an equivalent (B∗) image near B. An alternative possibility is that A∗ is not a gravita- tional image at all but is part of the emission from the lens- ing galaxy (or even a foreground object). In this scenario one has to attribute the proximity of A∗ to A, and the fact that the A∗ lies at the expected position angle for a lensed image, to coincidence. In the end, we find none of the possibilities satisfactory. It could be that we are deluted about the reality of sep- aration difference and there is nothing to explain but the different colours of the A and B images. 3.2 B0712+472 Radio observations revealed this system to be a four-image gravitational lens, although only the A, B and C images were visible in the optical WFPC2 images (Jackson et al. 1998). In the new NICMOS image (Fig. 2) we clearly see the D image also. We use the new data to revisit the question of the anomalous flux density ratios discussed in the earlier work. In the radio, optical and infrared bands, images A and C have the same flux density ratio (about 2.5:1) within the errors. The major discrepancy concerns the flux densities of B and D. The invisibility of D in V band and its marginal visibility in I can be ascribed to reddening in the lensing galaxy. Reddening was also initially thought to be the case for the large difference observed in the radio and optical flux density of component B (it has 80-90% of the flux of compo- nent A in the radio and only 30% of A in 555-nm). However, the fact that the B/A ratio remains constant throughout the optical and infrared within the errors while the inferred red- dening at 555 nm, AV ∼1 mag, requires the reddening to fall to < 0.2 mag at 1.6 µm which is clearly not the case 6 N. Jackson, E. Xanthopoulos, and I.W.A. Browne Figure 1. Top: HST optical (555-nm and 814-nm) images of B0218+357. Bottom: HST infrared (1.6 µm) images. North is up and East to the left. Each image is 3.′′69 on a side. For the WFPC2 images the scale is 45.5 mas/pixel and for the NICMOS images 43 mas/pixel. here as seen from the NICMOS images, argues against it. Variability is now also unlikely, since the infrared and opti- cal observations were taken over a year apart, much longer than the likely time delay between images, and yet the B/A optical/infrared flux density ratio has remained relatively constant (and very different from the radio flux density ra- tio). It seems therefore likely that we are, as suggested by Jackson et al. (1998), seeing an episode of microlensing. Fur- ther monitoring of this system over periods of years or tens of years (the typical timescale of microlensing in such a sys- tem) should reveal a gradual increase in the optical and infrared B/A flux density ratio. The positions of all objects are (just) consistent between the optical, infrared and radio images. The exception is an apparent shift of the position of the lensing galaxy at differ- ent wavelengths which is small but systematic with increas- ing wavelength. The lensed object appears to have a spectral index of α ∼ 0, where the flux F(λ)∝ λα in the rest-frame wave- length range of 240 nm -- 690 nm. This is much redder than the typical values of α ∼ −1 found for quasars (e.g. Jack- son & Browne 1991), implying that little quasar continuum is present. It has already been remarked that the object is severely underluminous (Jackson et al. 1998); it now appears that the continuum is unusually red for a quasar and is thus likely to have a substantial contribution from starlight, al- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 NICMOS images of JVAS/CLASS gravitational lens systems 7 Figure 2. Top: HST optical (555-nm and 814-nm) images of B0712+472. Bottom: HST infrared (1.6-µm) images, without and with subtraction of the point sources A,B,C,D. Note the arc that remains in the infrared NICMOS picture after subtraction of the point sources. Each image is 3.′′69 on a side. North is up and East to the left. For the WFPC2 images the scale is 45.5 mas/pixel and for the NICMOS images 43 mas/pixel. though some AGN component must be present to give the broad lines. This reinforces the conclusion that we may be dealing with an example of a mini-quasar inhabiting a host which is bright compared to the AGN component. Given the presumed magnifications (Jackson et al. 1998), the in- trinsic radio luminosity of log(L5GHz/W Hz−1sr−1)≃24.4 is only just in the radio-loud category as defined by Miller, Peacock & Mead (1990). The host galaxy of the lensed ob- ject is also visible, smeared into an arc in the NICMOS pic- ture (Fig. 2) and is considerably redder than the quasar at its centre; the arc is invisible in the optical pictures. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The optical-infrared colours of the lensing galaxy are consistent with the conclusion by Fassnacht & Cohen (1998) that the lens is an early-type galaxy. The 4000A break, at the galaxy's redshift of 0.41 (Fassnacht & Cohen 1998) occurs between 555 nm and 814 nm. 3.3 B1030+074 The discovery of the lensed system B1030+074 was reported by Xanthopoulos et al. (1998). The lensed images are of a quasar/BL Lac of redshift 1.535 (Fassnacht & Cohen 1998) 8 N. Jackson, E. Xanthopoulos, and I.W.A. Browne and are separated by 1.56 arcsec with a radio flux density ratio in the range 12 to 19. The lensing galaxy has a redshift of 0.599 and its spectrum is typical of an early type galaxy (Fassnacht & Cohen 1998). In Fig. 3 we show the WFPC2 V and I pictures (Xan- thopoulos et al., 1998), together with our new NICMOS 1.6µ picture. In all three bands the lensing galaxy is seen near the fainter B image. The peak of the galaxy light lies on the line joining the two images but there is also a secondary emission feature to the SE of the main part of the galaxy. Xanthopoulos et al. suggest that this secondary peak may be a spiral arm. The H-band data indicate that the colour of this feature is similar to that of the rest of the lensing galaxy, possibly suggesting that it is not a spiral arm but either part of the main galaxy or a companion object. The colours of the lensed images are consistent with our knowledge that the lensed object is a quasar or BL Lac object. Of interest, however, are the optical/infrared flux density ratios of the images which are considerably higher than any of those measured at radio frequencies. It is tempt- ing to attribute such differences to extinction but the fact that the H-band ratio is even larger than V and I ratios ar- gues against this. We suggest that the differences are best explained as arising from microlensing. Substructure in the lensing galaxy can also distort the radio flux ratios (as in mi- crolensing for optical fluxes), in particular for highly magni- fied images, without appreciably changing image positions. As Mao & Schneider (1998) note such substructure can be for example spiral arms in disk galaxies or structure caused by continuous merging and accretion of subclumps, as may possibly be the case for B1030+074, and can have a little effect on time delays and the determination of H0. 3.4 B1600+434 The lens system B1600+434 was discovered in the first phase of the CLASS survey (Jackson et al. 1995). It is a two-image system of separation 1392 mas; the lensing galaxy lies close to the southeastern component. In Fig. 4 we show the HST 555-nm and 814-nm images, together with the new NIC- MOS 1.6-µm image. The lensing galaxy is an edge-on spiral (Jaunsen & Hjorth 1997; Kochanek et al. 1999), which has been modelled in detail by Maller, Flores & Primack (1997) and by Koopmans et al. (1998). Photometry of this object is affected by the coincidence of the southeastern image, B, with the lensing galaxy. It is also likely that both optical images are variable (Jaunsen & Hjorth 1997). Jaunsen & Hjorth (1997) present ground- based BVRI photometry from which they deduce substantial reddening to be present in image B, almost certainly due to passage through the lensing galaxy. The infrared data lend support to this view, as the flux density ratios of the images in the infrared and radio are indistinguishable to within the errors. The unreddened image A has a spectral index α ∼ −1.7, which is roughly the spectral index of a normal quasar. 4 SUMMARY AND CONCLUSIONS Our NICMOS observations have proved successful at de- tecting both lensing galaxies and lensed objects. We find that often the flux density ratios of the images are differ- ent from those measured at radio wavelengths. Having in- frared colours enables us to distinguish between extinction and microlensing as explanations for the different flux den- sity ratios. We find evidence for both in different systems. Reddening is clearly affecting the B image in B1600+434 and may be part of the explanation for the puzzling sys- tem B0218+357. On the other hand, in both B0712+472 and B1030+074, the fact that the infrared and optical im- age flux density ratios are the same but different from the radio ones is strong evidence that microlensing is important in these objects. On a larger scale microlensing can have an important ef- fect on lensing statistics (see Bartelmann & Schneider 1990) and the fact the we see evidence in maybe three of the four systems examined here argues that the phenomenon is not as rare as previously thought. If microlensing is considered, single highly-magnified images can occur and for example the tight correlation between total magnification and flux ratio is weakened by microlensing. The arcs seen in B0712+472 illustrate another impor- tant fact. Many of the lensed objects have relatively sublu- minous AGN and in some cases the infrared emission from the host galaxy starlight dominates over that from the AGN. This is when arcs are seen. Other examples are B1938+666 (King et al. 1998), B2045+265 (Fassnacht et al. 1999) and B1933+503 (Marlow et al. 1999). With an arc the lensing galaxy potential is probed at many points and provides use- ful additional constraints for the lens mass models. ACKNOWLEDGMENTS This research was based on observations with the Hub- ble Space Telescope, obtained at the Space Telescope Sci- ence Institute, which is operated by Associated Universi- ties for Research in Astronomy, Inc., under NASA contract NAS5-26555. This research was supported by the European Commission, TMR Programme, Research Network Contract ERBFMRXCT96-0034 "CERES". We thank C. Kochanek for comments leading to an improvement in the first version of the paper. REFERENCES Bartelmann, M., Schneider, P, 1990, A&A, 239, 113 Bica, E., Alloin, D., 1987, A&AS, 70, 281 Biggs, A.D., Browne, I.W.A., Helbig, P., Koopmans, L. V. E., Wilkinson, P.N., Perley, R. A., 1998, MNRAS, 304, 349. Browne, I.W.A., Patnaik, A.R., Walsh, D., Wilkinson, P.N., 1993, MNRAS, 263, L32 Carilli, C.L., Rupen, M.P., Yanny, B., 1993, ApJ, 412, L59 Corbett, E.A., Browne., I.W.A., Wilkinson, P.N., & Patnaik, A.R., 1995. "Astrophysical Applications of Gravitational Lensing", ed. Kochanek, C., IAU Symposium 173, Reidel, Dordrecht Fassnacht, C.D., Cohen, J. G., 1998, AJ, 115, 377 Fassnacht, C.D., et al., 1999, AJ, 117, 658 Grundahl, F., Hjorth, J., 1995, MNRAS, 275, L67 Hjorth, J., 1997, in "Golden Lenses", workshop held at Jodrell Bank, June 1997, available at http://multivac.jb.man.ac.uk:8000/ ceres/workshop1/proceedings.html c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 NICMOS images of JVAS/CLASS gravitational lens systems 9 Figure 3. Top: HST optical (555-nm and 814-nm) images of B1030+074. Bottom: HST infrared (1.6-µm) images, raw (left) and with the two point objects A and B subtracted (right). The very bright A image makes a good PSF subtraction difficult. Each image is 3.′′69 on a side. North is up and East to the left. For the WFPC2 images the scale is 45.5 mas/pixel and for the NICMOS images 43 mas/pixel. Howarth, I. D., 1983, MNRAS, 203, 301 Jackson, N., Browne, I.W.A., 1991, MNRAS, 250, 414 Jackson, N. et al., 1995, MNRAS, 274, L25 Jackson, N. et al., 1998, MNRAS, 296, 483 Jaunsen, A. O., Hjorth, J., 1997, A&A, 317, L39 Keeton, C. R., Kochanek, C. S., 1997, ApJ, 487, 42 Keeton, C. R., Kochanek, C. S., Falco, E. E., 1998, ApJ, 509, 561 King, L. J., Jackson, N. N., Blandford, R. D., Bremer, M. N., Browne, I. W . A., de Bruyn, A. G., Fassnacht, C., Koopmans, L., Marlow, D. R., Wilkinson, P. N., 1998, MNRAS, 295, L41 Kochanek, C. S, 1996, ApJ, 466, 638 Kochanek, C. S, Falco, E. E., Impey, C. D., Leh´ar, J., McLeod, B. A., Rix H.-W., 1999, Proceedings of the 9th Annual Astro- physics Conference in Maryland, After the Dark Ages: When Galaxies Were Young, eds. S. Holt and E. Smith, American Institute of Physics Press, 1999, p. 163 Koopmans, L.V.E., de Bruyn, A.G., Jackson, N., 1998, MNRAS, 295, 534 Krist, J., 1997, "WFPC2 ghosts, scatter and PSF field depen- dence", postscript document available from the STScI WWW page Kundi´c, T., Turner, E.L., Colley, W.N., Gott, J.R., Rhoads, J.E., Wang, Y., Bergeron, L.E., Gloria, K.A., Long, D.C., Malho- tra, S., Wambsganss, J., 1997, ApJ, 482, L37 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 10 N. Jackson, E. Xanthopoulos, and I.W.A. Browne Figure 4. Top: HST optical (555-nm and 814-nm) images of B1600+434. Bottom: HST infrared (1.6-µm) image. Each image is 3.′′69 on a side. North is up and East to the left. For the WFPC2 images the scale is 45.5 mas/pixel and for the NICMOS images 43 mas/pixel. Larkin, J.E., Matthews, K., Lawrence, C.R., Graham, J.R., Har- rison, W., Jernigan, G., Lin, S., Nelson, J., Neugebauer, G., Smith, G., Soifer, B.T., Ziomkowski, C., 1994, ApJ, 420, L9 Lawrence, C.R., Cohen, J.G., Oke, J.B., 1995, AJ, 110, 2583 Lawrence, C.R., Elston, R., Januzzi, B.T., Turner, E.L., 1995, AJ, 110, 2570 Maller, A., Flores, R.A., Primack, J.R., 1997, ApJ, 486, 681 Mao, S., Schneider, P., 1998, MNRAS, 295, 587 Marlow, D.R., Browne, I.W.A, Jackson, N., Wilkinson, P. N., 1999, MNRAS, 305, 15 McLeod, B., Falco, E.E., Impey, C., Kochanek, C., Lehar, J., Rix, H.-W., Keeton, C., Munoz, J.A., Peng, C.Y., 1999, in preparation Miller, L., Peacock, J.A., Mead, A.R.G., MNRAS, 244, 207, 1990 Myers, S. T., et al., 1998, in preparation Patnaik, A.R., Browne, I.W.A., King, L.J., Muxlow, T.W.B., Walsh, D., Wilkinson, P.N., 1993, MNRAS, 261, 435 Patnaik, A.R., Porcas, R.W., Browne, I.W.A., 1995, MNRAS, 274, L5 Refsdal, S., Stabell, R., 1997, A&A, 325, 877 Saha, P., Williams, L. L. R., 1997, MNRAS, 292, 148 Schechter, P.L., et al., 1997, ApJ 475, L85 Stickel, M., Rieke, G.H., Kuhr, H., Rieke, M.J., 1996, ApJ, 468, 556 Urry, C. M., Padovani, P., 1995, PASP, 107, 803 Wiklind, T., Combes, F., 1995, A&A, 299, 382 Witt, H.-J., Mao, S., 1997, MNRAS, 291, 211 Xanthopoulos, E., et al., 1998, MNRAS, 300, 649 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
astro-ph/0309235
1
0309
2003-09-08T14:32:54
The extra-galactic Cepheid distance scale from LMC and Galactic period-luminosity relations
[ "astro-ph" ]
In this paper, we recalibrate the Cepheid distance to some nearby galaxies observed by the HST Key Project (KP) and the Sandage-Tammann-Saha (STS) group. We use much of the KP methodology in our analysis but apply new techniques, based on Fourier methods to estimate the mean magnitudes to published Cepheid data. We also apply different calibrating PL relations to estimate Cepheid distances, and investigate the sensitivity of the distance moduli to the adopted calibrating PL relation. We re-determine the OGLE LMC PL relations using a more conservative approach and also study the effect of using Galactic PL relations on the distance scale. For the KP galaxies, we find good agreement with an average discrepancy of -0.002 and 0.075mag. when using the LMC and Galaxy, respectively, as a calibrating PL relation. However, the Cepheid distance to STS galaxies are shorter by -0.178mag. and -0.108mag. on average again when using the LMC and Galaxy PL relations, respectively. We also calculate the distance to these galaxies using PL relations at maximum light and find very good agreement with mean light PL distances. However, after correcting for metallicity effects, the difference between the distance moduli obtained using the two sets of calibrating PL relations becomes negligible. This suggests that Cepheids in the LMC and Galaxy do follow different PL relations, at least at the median periods of the target galaxies.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: 04(11.04.1; 08.22.1) ASTRONOMY AND ASTROPHYSICS October 27, 2018 The extra-galactic Cepheid distance scale from LMC and Galactic period-luminosity relations S. M. Kanbur1, C. Ngeow1, S. Nikolaev2, N. R. Tanvir3, and M. A. Hendry4 1 Astronomy Department, University of Massachusetts, Amherst, MA 01003, USA 2 Institute for Geophysics and Planetary Physics, Lawrence Livermore National Laboratory, Livermore, CA 94550, USA 3 Department of Physical Science, University of Hertfordshire, College Lane, Hatfield, AL10 9AB, UK 4 Department of Physics and Astronomy, University of Glasgow, Glasgow G12 8QQ, UK Received 07 May 2003 / Accepted 14 Aug 2003 Abstract. In this paper, we recalibrate the Cepheid dis- tance to some nearby galaxies observed by the HST Key Project and the Sandage-Tammann-Saha group. We use much of the Key Project methodology in our analysis but apply new techniques, based on Fourier methods to esti- mate the mean of a sparsely sampled Cepheid light curve, to published extra-galactic Cepheid data. We also ap- ply different calibrating PL relations to estimate Cepheid distances, and investigate the sensitivity of the distance moduli to the adopted calibrating PL relation. We re- determine the OGLE LMC PL relations using a more conservative approach and also study the effect of using Galactic PL relations on the distance scale. For the Key Project galaxies after accounting for charge transfer effects, we find good agreement with an average discrepancy of -0.002 and 0.075mag. when us- ing the LMC and Galaxy, respectively, as a calibrat- ing PL relation. For NGC 4258 which has a geometric distance of 29.28mag., we find a distance modulus of 29.44 ± 0.06(random)mag., after correcting for metallic- ity. In addition we have calculated the Cepheid distance to 8 galaxies observed by the Sandage-Tammann-Saha group and find shorter distance moduli by −0.178mag. (mainly due to the use of different LMC PL relations) and −0.108mag. on average again when using the LMC and Galaxy, respectively, as a calibrating PL relation. However care must be taken to extrapolate these changed distances to changes in the resulting values of the Hubble constant because STS also use distances to NGC 3368 and 4414 and because STS calibration of SN Ia is often decoupled from the distance to the host galaxy through their use of differential extinction arguments. We also calculate the distance to all these galaxies using PL relations at maxi- Send offprint requests to: S. Kanbur email: [email protected] mum light and find very good agreement with mean light PL distances. However, after correcting for metallicity effects, the difference between the distance moduli obtained using the two sets of calibrating PL relations becomes negligible. This suggests that Cepheids in the LMC and Galaxy do follow different PL relations and constrains the sign for the coefficient of the metallicity correction, γ, to be neg- ative, at least at the median period log(P ) ≈ 1.4, of the target galaxies. Key words: Cepheids -- distance scale 1. Introduction The extra-galactic distance scale is one of the key prob- lems in modern astronomy. One of the basic parts of the solution is the correlation between period and mean lu- minosity (PL) obeyed by classical Cepheids. The Hubble Space Telescope (HST) H0 Key Project (Freedman et al. 2001, hereafter KP) has used Cepheids (discovered either by ground-based telescopes or by HST) in 31 spiral galax- ies, with 18 of them observed originally by KP, and the PL relation to estimate the distances to these galaxies. These distances were then used in turn to calibrate a host of sec- ondary distance indicators and hence estimate Hubble's constant. In parallel with this a team led by Sandage et al. (Saha et al. 2001b, hereafter STS) has used the HST to discover Cepheids in spiral galaxies which host Type Ia supernovae. Cepheid distances to these galaxies were used to calibrate the Hubble diagram for Type Ia supernovae and hence estimate Hubble's constant. Though both groups used mostly HST observations to discover the Cepheids and similar methodologies, the KP team favor a short distance scale and a larger value of 2 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations H0 (72± 8kms−1M pc−1, Freedman et al. 2001) whilst the STS group favor a long distance scale and smaller value of H0 (58.7 ± 6.3kms−1M pc−1, Saha et al. 2001b). The dis- crepancy in the value of the Hubble constant from these two groups is still unresolved (e.g., see Hendry & Kanbur 1996; Beaulieu et al. 1997; Kochanek 1997; Sasselov et al. 1997; Tanvir et al. 1999; Caputo et al. 2000; Gibson et al. 2000; Gibson & Stetson 2001). It is due only in part to the Cepheid distance scale. We note that the recent WMAP results (Spergel et al. 2003) find a Hubble constant very similar to the value obtained by KP. Nevertheless it is still very instructive to discover the reason behind the discrep- ancies in the Cepheid distance scale from the STS and KP groups. A proper understanding of the Cepheid PL relation in this regard is very important for an accurate local distance scale. In this paper we concentrate on re- calibrating the Cepheid distance to the target galaxies in both KP and STS groups with existing data because of the following factors: 1. HST Cepheid data is sparsely sampled with typically 12 and 4 points per Cepheid in the V and I band re- spectively (see Section 2 for details). In order to apply the Cepheid PL relations, it is necessary to estimate the mean magnitudes from these data in both V and I bands. In this paper, we use the techniques of Fourier expansion and interrelations to estimate the V and I mean magnitudes from sparsely populated light curves. These techniques are described in detail in Ngeow et al. (2003). 2. We re-analyze the OGLE (Optical Gravitational Lens- ing Experiment) LMC Cepheids (Udalski et al. 1999a) on the basis of the quality of their V band light curve, and develop more conservative LMC PL rela- tions based on this analysis and investigate the sensi- tivity of extra-galactic Cepheid distances to the LMC PL relation. This approach is different to the "sigma- clipping" methods currently used by Udalski et al. (1999a). 3. The average value of metallicities, defined as 12 + log(O/H), in target galaxies is about 8.84 ± 0.31dex (Freedman et al. 2001). This value is closer to the stan- dard Solar value of 8.87±0.07dex (Grevesse et al. 1996) than the LMC value of 8.50 ± 0.08dex (see reference in Ferrarese et al. 2000). Therefore, another approach is to use Galactic Cepheid PL relations as fundamen- tal calibrating relations (Feast 2001, 2003; Tammann 2003; Thim et al. 2003; Fouqu´e et al. 2003). We com- pare the distances obtained when using both the LMC and Galactic PL relations. 4. Kanbur & Hendry (1996) building on the hydrodynam- ical models of Simon et al. (1993) suggested that PL relations at maximum light may be more accurate than PL(mean) relations. To test this idea, we compute dis- tances using PL(max) relations and compare with their mean light counterparts. In Section 2 we summarize the photometric data used in this study. In Section 3 we present and discuss the methodology used in determining Cepheid distances, in- cluding obtaining the V and I band means and applying different PL relations. The results will be presented in Sec- tion 4. The conclusion and the discussion will be presented in Section 5. 2. The Photometric Data The photometry data for the Cepheids in each target galaxy were directly obtained from the corresponding pub- lished papers. We emphasize that we did not repeat the photometric reduction from raw data. The target galaxies that are selected in this study include: 16 KP galaxies, 8 STS galaxies and NGC 4258, which has an accurate geo- metrical distance measurement from water maser studies (Herrnstein et al. 1999, hereafter WM galaxy). The pho- tometric data of 16 KP galaxies were directly downloaded from the KP web-page1, excluding NGC 1425 (the data is not available at the time of analysis) and NGC 5457 (M101, as the observations to this galaxy include its outer field (Kelson et al. 1996) and inner field (Stetson et al. 1998b), which complicated the analysis). The photomet- ric data for the STS and WM galaxies were taken from the STS papers (see Saha et al. 2001b for references for each target galaxy)2 and Newman et al. (2001), respectively. The list of the target galaxies can be obtained from Table 4. There are two major photometry reduction pack- ages used in reducing the data for these galaxies, the ALLFRAME (Stetson 1994, 1996) and a variant of the DoPHOT (Schechter et al. 1993; Saha et al. 1996) pack- age. The KP team utilized both packages as a double- blind reduction process (Kennicutt et al. 1995; Freedman et al. 2001) to check the consistency between the two packages. However, the final, adopted distance moduli are based on ALLFRAME results (Freedman et al. 2001). Inter-comparison of the results from these two packages (galaxy-by-galaxy basis) show good agreement between the two (see the KP papers for more details and Hill et al. 1998). We also use the ALLFRAME photometric data downloaded from their web-page (if available). However, the photometric data for NGC 2541 and NGC 4321 are only available in DoPHOT from the same web-page. Sim- ilarly, only the DoPHOT photometric data is available for WM galaxy, although the photometry reductions were carried out by both packages (Newman et al. 2001). In contrast, the photometric data for the STS galax- ies were reduced mainly with DoPHOT, although some of 1 The URL is http://www.ipac.caltech.edu/H0kp/H0KeyProj .html. The reference to each KP galaxies also listed in this URL. 2 We exclude NGC 3368 (Tanvir et al. 1999) because it's not part of the STS program, and we are not calibrating the SN Ia distance in this paper. Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 3 (1) the galaxies also used an additional reduction package to check the DoPHOT results (e.g., reduction of NGC 4527 included the ROMAFOT package, see Saha et al. 2001a). Nevertheless, the available data for all STS galaxies are from the DoPHOT package. Although the KP team has reanalyzed the STS galaxies with the ALLFRAME pack- age (Gibson et al. 2000; Gibson & Stetson 2001; Stetson & Gibson 2001; Freedman et al. 2001), these data are not yet available from the KP web-page (as claimed in these papers). However, most, but not all, of the Cepheids are common in both STS and KP results (Gibson et al. 2000). Hence, we only can use the DoPHOT results in our anal- ysis of STS galaxies. The next steps in the photometric reduction process are a Cepheid search and period determination. Each re- duction package has its own algorithms to perform the period search, and the results generally agree well. We use published Cepheid data and periods in our analysis. The HST observations of the target galaxies generally consist of 24 V band and 8 I band cosmic-ray (CR) split images. The analysis of DoPHOT and ALLFRAME treat these images differently to deal with the cosmic-ray and point-spread functions. However, most of the published photometry combined the back-to-back CR split images together and tabulated the averaged 12 V band and 4 I band data points. Therefore, we only can use these (re- duced) photometric data points to reconstruct the light curves and obtain the mean magnitudes by the Fourier techniques described in the next section. 3. General Methodology and Analysis The physical basis of Cepheid PL relations and their us- age in determining the distance have been covered exten- sively in the literature (e.g., see Sandage & Tammann 1968; Feast & Walker 1987, Madore & Freedman 1991; Freedman et al. 2001) and would not be repeated here, only a brief description will be presented. The application of the Cepheid PL relation to estimate distances involves the discovery and appropriate observation of Cepheids suf- ficient to estimate their periods (in days) and mean mag- nitudes. In order to correct for the extinction/reddening, observations of extra-galactic Cepheids by HST are nor- mally taken in bandpasses V and I. The PL relation in bandpass λ can be expressed as: Mλ = aλ log(P ) + bλ, where a and b are the slope and zero-point, respectively. For the measurement of the apparent mean magnitudes of Cepheids in a target galaxy, < mλ >, the distance modu- lus in bandpass λ is: µλ =< mλ > −aλ log(P ) − bλ. Since µ0 = µV − AV = µI − AI , the reddening-free dis- tance modulus, or the Wessenheit function (Madore 1982; Moffett & Barnes 1986; Madore & Freedman 1991; Tanvir 1999), can be derived as: µ0 = µV − R(µV − µI ), where R ≡ AV /E(V − I) is the ratio of total-to- selective absorption. Following KP, we adopt R = 2.45 from Cardelli et al. (1989). Note that the validity of equa- tion (1) is based on the assumption that there are no corre- lated measurement errors in the V and I bandpasses. Then any differences in the V and I band distance moduli are due to differential reddening (see Saha et al. 1996, 2001b for details). After applying a period cut to the short period Cepheids in target galaxies (to avoid the incomplete- ness bias at the faint-end of observations, see Sandage 1988; Lanoix et al. 1999a; Freedman et al. 2001), an un- weighted mean of apparent distance moduli to the remain- ing Cepheids is adopted as the distance modulus to the target galaxy. The random (or statistical) error associated with equation (1) can be calculated via the standard for- mulae (i.e. σ/√N ). The reasons for taking the unweighted mean are: (a) the photometric errors of the mean magni- tudes are smaller than the expected width of the insta- bility strip, and are, together with other systematic er- rors, the dominant part of the weight in a weighted mean (Leonard et al. 2003); (b) this is equivalent to the fitting scheme used by KP; and (c) it can be easily incorporated with the weighting scheme adopted by the STS group (Tanvir 1997). Finally, due to the possible metallicity de- pendence of the Cepheid PL relation (see the discussions and reference in, e.g., Kennicutt et al. 1998; Freedman et al. 2001), metallicity corrections are added to equation (1): µz = µ0 + δz, (2) where δz = γ([O/H]ref −[O/H]gal) with the usual def- inition of [O/H] ≡ 12 + log(O/H). The reference metal- licity, [O/H]ref , is 8.50dex and 8.87dex for using LMC and Galactic PL relations, respectively. Some published values of γ range from −0.88 ± 0.16 and −0.56 ± 0.20 (Gould 1994), −0.44+0.10 −0.20 (Beaulieu et al. 1997; Sasselov et al. 1997), −0.4 ± 0.2 (Kochanek 1997), −0.32 ± 0.21 (Freedman & Madore 1990), −0.24±0.16 (Kennicutt et al. 1998) to 0.27 (Caputo et al. 2000). Note that the value of γ is method-dependent, i.e. it depends on the bandpasses and calibrating PL relation adopted. For example, strictly speaking the metallicity dependence quoted in Kennicutt et al. (1998) is only applicable when the Madore & Freed- man (1991) PL relation is used, and then only for the V and I bandpasses. Although there is some debate on the value for γ (see Freedman et al. 2001; Tammann et al. 2001 for details), we adopt γ ≡ γV,I = −0.2± 0.2mag dex−1 as in KP, who note that this value is in the middle of a num- ber of different determinations of γ. Changes in metallicity affect the mean brightness of a Cepheid and the papers 4 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations quoted above were aimed at quantifying this brightness shift due to metallicity. However, we now know that the Cepheid PL relation has a different slope in the Galaxy and LMC, and indeed may have two different slopes in the LMC itself. Thus it may be that metallicity affects not only the mean brightness of a Cepheid but also the way that mean brightness changes with period (the slope of the PL relation). We discuss this and its implications for metallicity corrections like equation (2) in Section 5. Since we are using the published photometry data to calculate the distance modulus, we share the same system- atic errors as in the published results. These include the uncertainty in the zero point of the PL relation, calibra- tion of photometric zero points, reddening and metallicity effects, incompleteness bias and crowding. The order of a typical systematic error is around ∼ 0.1 − 0.2mag. 3.1. The Mean Magnitudes As mentioned in Section 2, most of the HST observa- tions of extra-galactic Cepheids contain only 12 V band epochs and 4 (or 5) I band epochs. These observations use a power-law time series sampling strategy to minimize the aliasing problem and maximize the phase coverage of the observed data points within the observing windows (Freedman et al. 1994; Kennicutt et al. 1995). In order to reduce the bias due to sampling procedures, STS and some early KP observations used phase-weighted inten- sity means to find mean magnitudes in the V band. The I band mean magnitudes were found via empirical rela- tions developed by Freedman (1988) and Labhardt et al. (1997) for KP and STS observations, respectively (see, e.g., Silbermann et al. 1996). Other KP observations used a template-fitting procedure (Stetson 1996) to obtain the mean magnitudes in both V and I band simultaneously. In addition to the methods mentioned, Ngeow et al. (2003) recently developed an alternative method to ob- tain the mean magnitudes, which is based on Fourier tech- niques. Ngeow et al. (2003) give specific examples of sit- uations when such a method could be preferable to tem- plate based techniques. Another motivation for developing these Fourier techniques is the possibility of reconstruct- ing observed HST light curves to compare with models. In Section 4.1, we quantitatively compare the means ob- tained using our methods with existing techniques in the literature. Here, we apply the new Fourier method to all target galaxies in order to be consistent in doing the analysis, and also to test this new method in the distance scale problem. The details of using the Fourier techniques are presented in Ngeow et al. (2003). Here we outline the steps: 1. The V band: The light curves of 12 V band data points are reconstructed via a 4th-order Fourier expansion with the form: m(t) = m0 + i=4 X i=1 [Ai cos(2iπt/P + φi)] (3) where A and φ are Fourier amplitudes and phases, respectively, and m0 is the mean magnitude. Since the periods are directly taken from literature, the re- maining nine parameters can be obtained by fitting the data with equation (3). We use the technique of simulated annealing and restrict the ranges that the Fourier parameters in equation (3) can take to recon- struct the V band light curves. The ranges of Fourier phases are from 0 to 2π, whilst the ranges of Fourier amplitudes are determined from the "calibrating set" Cepheids and OGLE LMC Cepheids. The "calibrat- ing set" Cepheids consist of ∼100 Galactic Cepheids, mostly observed originally by Moffett & Barnes (1984), and some LMC/SMC Cepheids. 2. The I band: For 4 I band data points, it is clear that equation (3) cannot be applied to reconstruct the I band light curves. Therefore, we use Fourier interre- lations (Ngeow et al. 2003) to reconstruct the I band light curves. The Fourier interrelations are linear rela- tions connecting the Fourier parameters in the V and I bands: Ai(I) = αi + βiAi(V ); φi(I) = γi + ηiφi(V ) (4) The coefficients in equation (4) are determined from the "calibrating set" Cepheids. Ngeow et al. (2003) developed such Fourier interrelations for the Galaxy, LMC and SMC separately and found that they are only weakly dependent on metallicity. This is impor- tant since it means that the interrelations can be ap- plied over a wide range of metallicity. Given the so- lution of equation (4), the observed I band points are then used to establish the I band light curves and hence estimate the corresponding mean magnitudes by min- imizing the chi-square. 3. These Fourier techniques have been tested with Monte Carlo simulations, and show that the reconstruction procedures are unbiased and the errors of the Fourier amplitudes and means are around ∼ 0.03mag. Although the filters installed on the HST are close to the standard Johnson and Kron-Cousins system, i.e. F 555W ∼ V and F 814W ∼ I, conversions from the HST passbands to the standard photometry systems have been used by both KP and STS teams. For the case of the ALL- FRAME photometry reduction in the KP galaxies, the conversions are made during the reductions with the for- mulae suggested by Holtzman et al. (1995), hence the pub- lished photometry of the Cepheids are in standard bands. A similar situation regarding the conversion holds for the two KP galaxies with DoPHOT photometry (NGC 2541 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 5 Table 1. Mean V and I magnitudes from Fourier fits.a Galaxy-Cepheid NGC925-C5 NGC925-C6 NGC925-C7 log(P ) V(mag.) 1.686 23.680 24.533 1.635 1.624 24.305 I(mag.) 22.478 23.453 23.457 a The entire table is available electronically at the CDS. & NGC 4321) and the WM galaxy. For STS galaxies with DoPHOT photometric reductions, the conversions are ap- plied after the reductions to the calculated mean magni- tudes (see equations (5) and (6) in Saha et al. 1994 for WFPC; and equations (2) and (3) in Saha et al. 1996 for WFPC2). In the case of PL(max) relations, however, no such conversion is available yet. We have to assume the conversions at maximum light are similar to the equations given by Saha et al. (1994, 1996), and apply the conver- sion accordingly. These conversions have been applied to STS galaxies in Table 5, 6 & 7. The derived mean magnitudes in both V and I bands for the Cepheids used in this study are listed in Table 1, where column 1 shows the Cepheid in the target galaxy, column 2 is the period adopted from the literature, and columns 3 and 4 are the mean V and I band magnitudes, respectively, derived from the Fourier techniques described in this paper. Table 1 is available in its complete electronic form at the CDS3. Here we only show a portion of the table to indicate its form and content. 3.2. The Period-Luminosity Relations In order to apply the PL relations to HST data, we only look for the published PL relations that are available in both V and I bands in both the Galaxy and LMC. The following subsections discuss the adopted PL relations in this study. Note that for the LMC PL relations, we adopt µLM C = 18.50mag., to be consistent with the KP team (Freedman et al. 2001). All of the adopted PL relations are listed in Table 2 and 3, for the V and I bands, respectively. 3.2.1. The LMC PL Relations The LMC PL relations used by both KP team (before the publication of their final paper) and STS team are based on a homogeneous sample of 32 Cepheids, with periods ranging from 10 days to ∼120 days (Madore & Freedman 1991, hereafter MF91). Since this PL relation has been extensively used in determining Cepheid distances in the past, and the STS team still use this PL relation in their study (e.g., see Saha et al. 2001b), we adopt the MF91 re- lations as one of our calibrating PL relations in this study. After the publication of PL relations derived from OGLE LMC Cepheids (Udalski et al. 1999a, hereafter 3 http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/ U99), the KP team recalibrated their Cepheid distances with these new LMC PL relations in their final paper, as well as the Cepheid distance to NGC 4258 by Newman et al. (2001). These reddening-corrected PL relations were derived using Cepheids with log P > 0.4 to minimize pos- sible contamination by first overtone pulsators and were "sigma clipped", resulting in ∼ 650 Cepheids used in de- riving the PL relations. In order to compare our results with published Cepheid distances, we also adopt the U99 PL relations. The new U99 PL relations have dramatically changed Cepheid distances compared to distances derived from MF91 PL relations, in the sense that the derived dis- tances are smaller with U99 PL relations (Freedman et al. 2001). There are some criticisms about the U99 PL relations in the literature. The U99 PL relations are dominated by short period Cepheids (with < log(P ) >∼ 0.5 and about 90% of them have period shorter than 10 days), and a lack of longer period Cepheids with log(P ) > 1.5 (Feast 2001, 2003; Saha et al. 2001b). However, Freedman et al. (2001), by using Sebo et al. (2002) LMC data, claimed that the use of the OGLE LMC PL relations in estimat- ing distances to target galaxies whose Cepheids all had periods longer than the longest period Cepheid observed by Udalski et al. (1999a) made little difference. Another potential problem of the U99 PL relations is the discovery of a break in the PL relation at log(P ) = 1.0, as shown in Tammann et al. (2001) and in Tammann & Reindl (2002). The long and short period Cepheids follow different PL relations. The reasons for the break in PL relations are still unclear, nevertheless we adopt the PL relations for long period Cepheids (Tammann & Reindl 2002, hereafter TR02) since most of the extra-galactic Cepheids have pe- riods longer than 10 days. Despite these potential problems, we re-analyze OGLE LMC Cepheid data and re-fit the PL relations without us- ing sigma-clipping. The sigma-clipping method is an iter- ative procedure whereby the data are fitted with a regres- sion and then those points lying 2.5σ away from the regres- sion line are removed. A new regression is fitted and the procedure continues for few cycles. Therefore, it blindly removes all the outliers to reduce the scatter in the fi- nal fit. However, we feel this approach may both remove some true Cepheids from the data and include some suspi- cious Cepheids in the final sample (see discussion below). In addition, Nikolaev et al. (2003) discussed some flaws associated with the sigma-clipping algorithm, including the implicit assumption of the normal distribution of the residuals and the sensitivity of the results to the chosen threshold of kσ. We use the same OGLE LMC data as in Udalski et al. (1999a) to derive the LMC PL relations, but with- out processing the sigma-clipping procedure (Udalski et al. 1999a; Willick & Batra 2000; Groenewegen 2000). Instead, we examine in detail the light curves for all Cepheids. The OGLE photometric data were downloaded 6 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations Table 2. The adopted V band Period-Luminosity relations.a Relation LMC (MF91) OGLE LMC (U99) OGLE LMC (Here) LMC PL>10d (TR02) LMC PL(Max) Galactic (GFG98) Galactic (FSG03) Galactic (T03) Slope (aV ) −2.76 ± 0.11 ZP (bV ) −1.40 ± 0.05 −2.760 ± 0.031 −1.458 ± 0.021 −2.746 ± 0.043 −1.401 ± 0.030 −2.48 ± 0.17 −1.75 ± 0.20 −2.744 ± 0.051 −1.817 ± 0.035 −3.037 ± 0.138 −1.021 ± 0.040 −0.989 ± 0.034 −3.141 ± 0.100 −0.826 ± 0.119 −3.06 ± 0.11 σV 0.27 0.159 0.223 0.16 0.261 0.209 · · · 0.24 a For LMC PL relations, Assume µLM C = 18.50mag. Table 3. The adopted I band Period-Luminosity relations.a Relation LMC (MF91) OGLE LMC (U99) OGLE LMC (Here) LMC PL>10d (TR02) LMC PL(Max) Galactic (GFG98) Galactic (FSG03) Galactic (T03) Slope (aI) ZP (bI ) −3.06 ± 0.07 −1.81 ± 0.03 −2.962 ± 0.021 −1.942 ± 0.014 −2.965 ± 0.028 −1.889 ± 0.019 −2.82 ± 0.13 −2.09 ± 0.15 −2.958 ± 0.033 −2.129 ± 0.023 −3.329 ± 0.132 −1.435 ± 0.037 −1.550 ± 0.034 −3.408 ± 0.095 −1.325 ± 0.114 −3.24 ± 0.11 σI 0.18 0.109 0.145 0.12 0.171 0.194 · · · 0.23 a For LMC PL relations, Assume µLM C = 18.50mag. N 32 649 634 ∼47 634 28 32 53 N 32 658 634 ∼47 634 27 32 53 15 15.5 16 16.5 17 17.5 0 14 14.5 15 15.5 16 16.5 0 C164, log(P)=0.501 0.2 0.4 0.6 0.8 1 phase 0.2 0.4 0.6 0.8 1 phase 14.5 15 15.5 16 16.5 17 0 14 14.5 15 15.5 16 16.5 0 C207, log(P)=0.754 0.2 0.4 0.6 0.8 1 phase 0.2 0.4 0.6 0.8 1 phase Fig. 1. Two examples of unusual light curves in OGLE Cepheids. Upper panels are V band light curve and lower panels are I band light curves. The light curves for Cepheids that have similar periods are shown with dashed lines, for comparison. Original data points are also indicated. from the OGLE web-page4. This consisted of 771 funda- mental mode Cepheids (as judged by Udalski et al. 1999b). Then, short period Cepheids (log(P ) < 0.4) were elimi- nated from the sample, as in the case of Udalski et al. 4 The URL is http://bulge.princeton.edu/∼ogle/ (1999a). The V and I band photometric data for the re- maining Cepheids were then fit with the 4th-order Fourier expansion using simulated annealing techniques (equation (3)) to reconstruct the light curves and obtain the mean magnitudes (Ngeow et al. 2003). The light curves for each Cepheid were then visually inspected, and further elimi- Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 7 15 15.5 16 16.5 17 0 15 15.5 16 16.5 17 0 C486, log(P)=0.515 dip 0.2 0.4 0.6 0.8 1 phase C617, log(P)=0.520 bump gap 0.2 0.4 0.6 0.8 1 phase 15 15.5 16 16.5 17 0 15 15.5 16 16.5 17 0 C761, log(P)=0.477 clustering of data 0.2 0.4 0.6 0.8 1 phase C770, log(P)=0.501 scattering of data 0.2 0.4 0.6 0.8 1 phase Fig. 2. Some examples of poorly constructed V band light curves, due to bad phase coverage or clustering of data points in certain phases. Original data points are also indicated. 15 15.5 16 16.5 17 0 13.5 14 14.5 15 15.5 0 C596, log(P)=0.762 0.2 0.4 0.6 0.8 1 phase 0.2 0.4 0.6 0.8 1 phase 14.5 15 15.5 16 16.5 0 14 14.5 15 15.5 16 0 C522, log(P)=0.655 0.2 0.4 0.6 0.8 1 phase 0.2 0.4 0.6 0.8 1 phase Fig. 3. Some examples of Cepheids that have unusually small amplitudes. The light curves for the Cepheids that have similar periods are shown in dashed lines, for comparison. Original data points are also indicated. nation of data was carried out according to the following criteria: 1. In order to have a self-consistent number of Cepheids in both bands, Cepheids without V band data were removed from the sample. 2. There are 3 Cepheids that have very unusual light curves, with two of them shown in Figure 1. The nature of these objects or the reason for having the unusual light curves is still unknown. However it is clear that they must be eliminated from the sample. 3. Then, we visually examined the V band data and cor- responding Fourier expansions for each star to deter- mine which stars had acceptable Fourier expansions. In some cases the V band data are clustered around a phase point resulting in numerical bumps in the Fourier expansion (see Ngeow et al. 2003 for more details). This is never a problem in the I band be- cause the number of points per light curve is so much greater (more than 120, as compare to V band which has ∼ 12-∼ 50 data points per light curves). However, the Cepheids with poorly reconstructed V band light curves were eliminated in order to have exactly the same Cepheids in both bands. The number of stars re- 8 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 15 15.5 16 16.5 17 0 14.5 15 15.5 16 0 C1, log(P)=0.531 0.2 0.4 0.6 0.8 1 phase 0.2 0.4 0.6 0.8 1 phase 14 14.5 15 15.5 16 0 13 13.5 14 14.5 15 0 C649, log(P)=0.918 0.2 0.4 0.6 0.8 1 phase 0.2 0.4 0.6 0.8 1 phase Fig. 4. Two examples of well reconstructed light curves. The upper panels are V band light curves and the lower panels are I band light curves. Original data points are also indicated. jected according to this criterion is 34. Some examples of this case are shown in Figure 2. than the (classical) Type I Cepheids.5 Seven outliers are removed according to this criterion. 4. Finally, some Cepheids appear to have a flat or low amplitude light curves when compared to light curves of stars with similar periods. Figure 3 shows a few ex- amples of this case. Since we eliminated the Cepheids with log(P ) < 0.4, these variables are unlikely to be the RR Lyrae-type variables. Due to the uncertainty of their classification, we eliminated all 43 variables falling into this category. We further discuss this group of Cepheids in Section 4.4. We fit the PL relations to the remaining sample (two examples are shown in Figure 4) with the published ex- tinction values given by OGLE team, assuming AV = 3.24E(B − V ) and AI = 1.96E(B − V ) (Udalski et al. 1999b), to obtain the dereddened PL relations. The ex- tinction map of OGLE LMC Cepheids is derived from observations of red clump stars along 84 lines-of-sight to- wards the LMC (Udalski et al. 1999b). Although there are some criticisms of the OGLE extinction map (e.g., see Feast 2001; Beaulieu et al. 2001; Fouqu´e 2003), we adopt the same extinction map to be consistent with the OGLE team. After fitting the PL relations, there are a few out- liers showing up in the plots (not shown) that are more than ∼ 4σ away from the ridge lines at given periods. If we assume a normal distribution, these outliers can be re- jected (as in the case of sigma-clipping). However, these outliers are systematically fainter than other Cepheids at similar periods. For example, the V band magnitudes of these outliers are about 1.0mag fainter than the mean magnitudes predicted from the ridge lines. Hence, there is a suspicion that these outliers could be the Type II Cepheids (or W Virginis), as they are generally fainter The resulting PL relations with this final sample are presented in Table 2 and 3, and the corresponding plots are given in Figure 5(a). In addition, we did not fit the PL relations to long period Cepheids, as in Tammann et al. (2001), although it can be done easily. In Figure 5(b), we also plot out the positions of the eliminated Cepheids in PL plots. As can be seen from the figure, most of the out- liers are rejected, as with sigma-clipping. However, some eliminated Cepheids fall along the ridge lines and would not be rejected by the sigma-clipping method. Our philos- ophy is that it is better to select bona fide Cepheids, and remove dubious Cepheids that can be eliminated based on physical grounds. Further investigation of the eliminated Cepheids, e.g. the low amplitude Cepheids or the three Cepheids with unusual light curves, are outside the scope of this paper. It is true that in the end there is little differ- ence in our V and I band PL slope relation to that initially published by the OGLE team and used by the KP. Never- theless, in the age of precision cosmology it is important to use appropriate data. Further our OGLE sample will be important when comparing with pulsation model results. Moreover, with mounting evidence that the PL relation in the LMC is "broken", it is more appropriate to fit two PL relations to the LMC data. However we refit the OGLE data with one relation and included MF91 in our analy- sis to thoroughly test the effect of using our new way of 5 The sigma-clipping algorithm will automatically remove these outliers, but there is no physical reasoning for doing that. In addition, the mean magnitudes for these stars have been cor- rected for extinction of AV ∼ 0.5mag, unless these stars suffer higher extinction than the OGLE extinction map in order to produce the faint magnitudes. Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 9 13 14 15 16 17 12 13 14 15 16 0.4 0.6 0.8 1 log(P) 1.2 1.4 13 14 15 16 17 12 13 14 15 16 0.4 0.6 0.8 1 log(P) 1.2 1.4 0.4 0.6 0.8 1 log(P) 1.2 1.4 0.4 0.6 0.8 1 log(P) 1.2 1.4 Fig. 5. Left : (a) The V and I band PL relations from the final OGLE LMC sample, after correction for reddening. The best fit PL relations are indicated in solid lines. The dashed lines show the 2σ dispersion, as expected from the width of instability strip. Right : (b) The locations of eliminated Cepheids in PL relations. The symbols are: filled circles = 3 Cepheids with unusual light curves (Figure 1); open circles = 34 Cepheids with poorly reconstructed light curves (Figure 2); filled squares = 43 Cepheids with low amplitudes (Figure 3) and open triangles = 7 Cepheids that appear dimmer (by ∼ 1.0mag in V band) from the ridge line at given periods. calculating V and I band means. This is our motivation for studying T02 (the LMC PL relation only for LMC Cepheids with periods longer than 10 days). It must also be pointed out that the calibrating Cepheids in the LMC and Galaxy have periods less than 25 days but many of the extra-galactic Cepheids observed by HST have peri- ods longer than 25 days. This is common to our work and that of KP and STS. 3.2.2. The Galactic PL Relations Due to high extinction and less accurately known dis- tances to Galactic Cepheids from other independent meth- ods, Galactic PL relations have been considered less fa- vorably in the extra-galactic distance scale problem than LMC PL relations (e.g., Kennicutt et al. 1995). However, the idea of using Galactic PL relations has been revived recently (Feast 2001, 2003; Tammann 2003; Fouqu´e et al. 2003; Thim et al. 2003) because not only is the metallic- ity in many of the HST observed target galaxies closer to Galactic values, but also the current calibration of Galac- tic PL relations has recently been improved (e.g., see Feast 2003; Fouqu´e et al. 2003). In addition, because of some possible problems associated with the OGLE LMC PL re- lations, including the break at 10 days and the dominance of short period Cepheids (see previous subsection), and the possible small "depth effect" of LMC Cepheids (Groe- newegen 2000; Nikolaev et al. 2003), using the Galactic PL relations to calibrate Cepheid distances could be more desirable and can certainly provide an independent check of the LMC based scale which is independent of the LMC distance modulus: this is currently the largest source of systematic uncertainty in the extra-galactic distance scale (±0.1mag). This does not imply that the systematic un- certainly associated with Galactic PL relations is smaller, but if this is true, the determination of Cepheid distances can be improved. We pick three recent Galactic PL relations which have both V and I band PL relations. The first Galactic PL re- lation is derived from 28 Cepheids using the Barnes-Evans surface brightness technique (Gieren et al. 1998, hereafter GFG98). The updated version of GFG98 PL relations has become available just recently. This includes 32 Galactic Cepheids in their sample (Fouqu´e et al. 2003, hereafter FSG03). The last Galactic PL relation is adopted from Tammann et al. (2003), and includes the 28 Cepheids in GFG98 sample with an additional 25 Cepheids from Feast (1999), referred as T03 in Table 2 and 3. We did not include the Galactic PL relations resulting from the Hipparcos calibration because these PL relations assumed the V band slopes are the same as in LMC PL relations, and hence calibrate the zero-point only (e.g., see Feast & Catchpole 1997; Lanoix et al. 1999b; Groenewegen & Oudmaijer 2000; Feast 2003). 3.2.3. LMC PL(Max) Relations Kanbur & Hendry (1996), on the basis of hydrodynamical pulsation calculations (Simon et al. 1993), suggested that PL relations at maximum light may have smaller scat- 10 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations Table 4. The CTE and metallicity corrections (see text for details). Galaxy Nceph [O/H]a δz(LM C)b δz(GAL)b δCT E NGC 925 NGC 1326A NGC 1365 NGC 2090 NGC 2541 NGC 3031 NGC 3198 NGC 3319 NGC 3351 NGC 3621 NGC 4321 NGC 4414 NGC 4535 NGC 4548 NGC 4725 NGC 7331 IC 4182 NGC 3627 NGC 3982 NGC 4496A NGC 4527 NGC 4536 NGC 4639 NGC 5253 NGC 4258 72 15 47 30 29 17 36 33 48 59 42 8 47 24 15 13 27 25 14 45 13 31 15 5 15 KP Galaxies 8.55 ± 0.15 8.50 ± 0.15 8.96 ± 0.20 8.80 ± 0.15 8.50 ± 0.15 8.75 ± 0.15 8.60 ± 0.15 8.38 ± 0.15 9.24 ± 0.20 8.75 ± 0.15 9.13 ± 0.20 9.20 ± 0.15 9.20 ± 0.15 9.34 ± 0.15 8.92 ± 0.15 8.67 ± 0.15 STS Galaxies 8.40 ± 0.20 9.25 ± 0.15 (8.9 ± 0.4) 8.77 ± 0.15 (8.9 ± 0.4) 8.85 ± 0.15 9.00 ± 0.15 8.15 ± 0.15 WM Galaxies 8.85 ± 0.15 0.010 0.0 0.092 0.060 0.0 0.050 0.020 -0.024 0.148 0.050 0.126 0.140 0.140 0.168 0.084 0.034 -0.020 0.150 0.080 0.054 0.080 0.070 0.100 -0.070 -0.064 -0.074 0.018 -0.014 -0.074 -0.024 -0.054 -0.098 0.074 -0.024 0.052 0.066 0.066 0.094 0.010 -0.040 -0.094 0.076 0.006 -0.020 0.060 -0.004 0.026 -0.144 0.070 -0.004 -0.07 -0.07 -0.07 -0.07 -0.07 0.0c -0.07 -0.07 -0.07 -0.07 -0.07 -0.07 -0.07 -0.07 -0.07 -0.07 0.0c 0.05 0.05 0.05 0.05 0.05 0.05 0.0c 0.0c a [O/H] ≡ 12 + log(O/H), adopted from Ferrarese et al. (2000), except for NGC 3982 (Stetson & Gibson 2001) and NGC 4527 (Gibson & Stetson 2001). b δz = (−0.2 ± 0.2) mag dex−1([O/H]ref − [O/H]). For LMC, [O/H]ref = 8.50dex and for Galactic, [O/H]ref = 8.87dex. c Since the observations of NGC 3031, IC 4182 and NGC 5253 were done with WFPC (not WFPC2), CTE correction is not included. The reduction of NGC 4258 used Stetson's (1998) calibration (Newman et al. 2001), no CTE correction is needed. ter. However, they found that for a sample of 32 stars in the LMC, the PL relation at maximum light (here- after PL(Max)) in the LMC had comparable scatter to its counterpart at mean light. Motivated by this, and to use it primarily as a check on our mean light results, we computed the distances to all target galaxies studied in this paper using PL(Max) relations. Maximum light for Cepheids in the target galaxies are estimated using the light curve reconstruction techniques described in Ngeow et al. (2003). We use the same OGLE LMC Cepheid sample as in the previous subsection to derive the PL(Max) relations. The estimation of maximum light (or the equivalent mini- mum magnitude) of a Cepheid from its light curve is more severely influenced by the quality of reconstructed light curves, as compared to its mean light counterpart. This is because bad phase coverage in the data will result in nu- merical bumps in the reconstructed light curves (Ngeow et al. 2003) which can be higher than the maximum light. Therefore, the elimination of poorly reconstructed light curves is more important in this aspect. After the maxi- mum light of each Cepheid in the sample has been esti- mated from the reconstructed light curves, the PL(Max) relations can be obtained. For this work, we assume that the reddening at maximum light is as same as at mean light even though the period color relations at maximum light are different to those at mean light (Simon et al. 1993). 3.3. CTE and Metallicity Corrections The charge-transfer efficiency (CTE) for WFPC2 is a com- plicated issue. Simply speaking, it has been found that the performance of the WFPC2 depends on the exposure time, which lead to the "long-vs-short" exposure correc- tions, and other factors (see Stetson 1998; Freedman et al. 2001 for more details). The published photometry data for KP galaxies were based on the Hill et al. (1998) calibra- Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 11 tion, but the final results presented in Freedman et al. (2001) are in the Stetson (1998) calibration. To convert the calibration from the Hill system to the Stetson sys- tem, a correction of δCT E = −0.07± 0.02mag is added to the distance modulus (Mould et al. 2000; Freedman et al. 2001). In contrast, the STS galaxies adopted the Holtz- man et al. (1995) calibration, which leads to a correction of δCT E = 0.05mag (see Saha et al. 2001a for example). The calibration used in the WM galaxy is in the Stetson system (Newman et al. 2001), hence no CTE correction is needed. Note that no CTE correction is added to the three galaxies (IC 4182, NGC 3031 & NGC 5253) observed with WFPC in the early days of HST. The CTE corrections for all target galaxies are summarized in the last column of Table 4, and the CTE corrected distance modulus will be represented as µ0,CT E. Due to metallicity differences between target galaxies and the LMC, metallicity corrections, as given in equa- tion (2), are commonly applied to µ0,CT E (Freedman et al. 2001). The metallicity of the target galaxies, in terms of [O/H], is given in column 3 of Table 4, and the corre- sponding metallicity corrections with respect to the LMC, δz(LM C), are tabulated in column 4 for the same table. Although the (average) metallicity of the target galaxies is closer to the Galactic value than the LMC (hence the metallicity correction is small), there is no obvious reason for not applying the same metallicity correction when us- ing a Galactic calibrating PL relation. Therefore, we apply the same correction, δz(GAL), to the distance moduli as in the case of using LMC PL relations (thought see Section 4.4). These metallicity corrections are listed in column 5 of Table 4 for each target galaxies. 4. Results We use the same Cepheids in each target galaxy that were used by the original KP or STS study after various selec- tion criteria (including the period-cut, color-cut and the quality of the Cepheids) had been applied. The periods of the Cepheids in final samples were taken from the cor- responding papers. The mean magnitudes were obtained with the Fourier techniques described in Section 3.1. Then the reddening-corrected distance modulus to individual Cepheids was calculated from equation (1) by using ei- ther the LMC or Galactic PL relations. The unweighted mean of the distance moduli to individual Cepheids was taken to be the final distance modulus to the target galaxy. The results of the distance moduli are presented in Table 5 and 6 when using both the LMC and Galactic PL relations respectively. The results obtained from LMC PL(Max) re- lations are given in Table 7. Note that the distance moduli in these tables have been corrected for CTE (column 6 in Table 4) but not for metallicity effects, and all the errors in the tables are random (statistical) errors only. 4.1. Comparisons of the Mean Magnitudes We compare the difference of mean magnitudes obtained with our method to published values. The results of this comparison are presented in Table 9, where we give the av- erage difference of mean magnitudes (∆/N ) in both bands for each target galaxy. Overall, the differences for the 16 KP galaxies are: < ∆V >= 0.039 ± 0.012mag. and < ∆I >= 0.032 ± 0.009mag.; for 8 STS galaxies: < ∆V >= 0.035± 0.010mag. and < ∆I >= −0.021± 0.008mag.; and for all 25 target galaxies: < ∆V >= 0.036±0.008mag. and < ∆I >= 0.014 ± 0.008mag. We also compare our distance moduli, calculated us- ing our Fourier techniques to estimate the V and I band means, with published distance moduli. The only two PL relations we can use in this comparison are the LMC PL relations from MF91 and U99, for STS and KP+WM galaxies, respectively. The results of these comparisons are presented in Table 10, which shows the average difference between our distance moduli and the published values. In overall, the difference is small among all the target galax- ies, with a difference of −0.020±0.017mag. (corresponding to ∼1% change in distance). This result implies that our method of calculating means is a viable alternative tech- nique, as can be seen from Figure 6. Ngeow et al. (2003) show that our method has advantages in some situations. The largest and smallest difference comes from NGC 4639 and NGC 4414, with a difference of −0.240 ± 0.245mag. and 0.004 ± 0.066mag., respectively. While STS use phase weighted averages to compute their means, the KP use light curve template techniques (Stetson 1996). The difference between this and the method adopted here is described in detail in Ngeow et al. (2003). Both methods fit equations like equation (3) to the observed data. In the case of Stetson (1996), A1 is ob- tained from the data while Ak, k > 1 is obtained from the templates showing how the ratio of Ak/A1, k > 1 varies with period. In our case all four parameters Ak, k ≥ 1 are obtained from the data with a simulated annealing approach. We do this, firstly, because we want to recon- struct the light curve to model it and, secondly, to capture any light curve shape changes due to metallicity. The top two panels of Figure 7 show the second Fourier amplitude coefficient (A2) calculated with this method for very well sampled Galactic (left panel) and LMC (right panel)data. Since this is very well sampled data, it is safe to assume that the progression of A2 with period is a good represen- tation of reality. The bottom panel shows the same Fourier amplitude calculated using the Stetson (1996) method. It can clearly be seen that at all periods the range of A2 at given period is larger than would be predicted by the Stet- son (1996) technique. Furthermore, from the well sampled data, we randomly pick 12 points (so that they could be clustered or not as the case may be), add Gaussian errors to these points and then calculate a simulated annealing 12 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations Table 5. The Cepheid distance (with CTE correction) to nearby galaxies with different LMC PL relations. MF91a Galaxy 29.801 ± 0.060 NGC 925 NGC 1326A 31.102 ± 0.088 NGC 1365 31.315 ± 0.052 NGC 2090 30.424 ± 0.036 NGC 2541 30.462 ± 0.072 27.887 ± 0.094 NGC 3031 NGC 3198 30.837 ± 0.060 NGC 3319 30.593 ± 0.094 NGC 3351 29.979 ± 0.090 NGC 3621 29.259 ± 0.059 30.900 ± 0.067 NGC 4321 NGC 4414 31.236 ± 0.052 NGC 4535 31.001 ± 0.054 30.867 ± 0.076 NGC 4548 NGC 4725 30.522 ± 0.066 NGC 7331 30.891 ± 0.083 28.365 ± 0.128 IC 4182 30.238 ± 0.083 NGC 3627 NGC 3982 31.714 ± 0.136 NGC 4496A 30.959 ± 0.043 30.755 ± 0.129 NGC 4527 NGC 4536 30.961 ± 0.102 NGC 4639 31.790 ± 0.108 27.881 ± 0.244 NGC 5253 29.435 ± 0.058 NGC 4258 U99a Herea TR02a 29.730 ± 0.060 31.007 ± 0.088 31.191 ± 0.051 30.328 ± 0.036 30.355 ± 0.071 27.800 ± 0.093 30.728 ± 0.060 30.511 ± 0.092 29.914 ± 0.090 29.164 ± 0.059 30.769 ± 0.067 31.104 ± 0.043 30.879 ± 0.053 30.783 ± 0.078 30.399 ± 0.067 30.800 ± 0.078 28.334 ± 0.130 30.134 ± 0.084 31.599 ± 0.134 30.839 ± 0.043 30.635 ± 0.127 30.830 ± 0.102 31.664 ± 0.105 27.873 ± 0.246 29.384 ± 0.056 29.718 ± 0.060 30.998 ± 0.088 31.186 ± 0.051 30.319 ± 0.036 30.348 ± 0.071 27.791 ± 0.093 30.721 ± 0.060 30.501 ± 0.092 29.902 ± 0.090 29.155 ± 0.059 30.764 ± 0.067 31.099 ± 0.044 30.874 ± 0.053 30.773 ± 0.078 30.393 ± 0.067 30.791 ± 0.079 28.318 ± 0.130 30.127 ± 0.084 31.593 ± 0.134 30.833 ± 0.043 30.629 ± 0.128 30.826 ± 0.102 31.659 ± 0.106 27.854 ± 0.246 29.370 ± 0.056 29.744 ± 0.060 31.027 ± 0.088 31.218 ± 0.052 30.348 ± 0.036 30.378 ± 0.071 27.818 ± 0.093 30.751 ± 0.060 30.528 ± 0.093 29.927 ± 0.090 29.184 ± 0.059 30.798 ± 0.067 31.133 ± 0.045 30.906 ± 0.053 30.800 ± 0.078 30.426 ± 0.067 30.819 ± 0.079 28.339 ± 0.129b 30.157 ± 0.083 31.624 ± 0.135 30.865 ± 0.043 30.661 ± 0.128 30.859 ± 0.102 31.692 ± 0.106 27.872 ± 0.245c 29.393 ± 0.057 a MF91 = Madore & Freedman (1991) PL relations; U99 = Udalski et al. (1999a) PL relations; Here = PL relations derived in Section 3.2.1; TR02= Tammann & Reindl (2002) PL relations for log(P ) > 1.0. See Section 3.2.1 for details. b There are 12 Cepheids in this galaxy with period less than 10 days. If we use the PL<10d relations, as given by Tammann & Reindl (2002), then µ = 28.340 ± 0.130, which is identical to the value with PL>10d relations. c There are 2 Cepheids in this galaxy with period less than 10 days. If we use the PL<10d relations, as given by Tammann & Reindl (2002), then µ = 27.876 ± 0.246, which is very close to the value with PL>10d relations. fit, we recover trends in the top panel of Figure 7. This is clearly displayed in figure 18 of Ngeow et al. (2003). 4.2. Distance Moduli from Various PL Relations Since the distance moduli listed in Table 5 are calculated using different LMC PL relations, they can be compared. A quick comparison in Table 5 indicates that the MF91 PL relation always give the longer distance moduli, and the shortest distance moduli are from the OGLE LMC PL relations derived in Section 3.2.1. The average difference between the distance moduli derived from MF91 and the OGLE LMC (Here), for all target galaxies in Table 5, is about 0.105 ± 0.006mag. Comparing this result with the previous case (i.e. Section 4.1) of using the same PL rela- tions but different means magnitudes, the distance modu- lus is more sensitive to different calibrating PL relations. Also, the distance moduli from three OGLE LMC PL rela- tions are consistent with each other within the statistical errors. A similar situation exists when using Galactic PL rela- tions, as shown in Table 6. The distance moduli from both GFG98 and FSG03 Galactic PL relations are consistent with each other (with a difference of ∼ 0.024± 0.007mag., as there are 26 common Cepheids in both samples), while FSG03 PL relations produce a shorter distance. However, the distance moduli from T03 PL relations are systemat- ically further than the other two, although all three PL relations share some common Galactic Cepheids. Never- theless, the average difference between the distance moduli from T03 and FSG03 is ∼ 0.097± 0.008mag., comparable to the difference seen in LMC PL relations. Finally, we compare the distance moduli (after CTE corrections) from LMC PL(Max) relations in Table 7 to their mean light counterparts, i.e., column 4 of Table 5. The distance moduli from PL(Mean) and PL(Max) rela- tions are consistent with each other, although PL(Max) relations generally give a shorter distance. This serves as an important check on our calibration of the extra- galactic distance scale. The closer distance from PL(Max) relations is expected (with the difference of ∼ 0.015 ± Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 13 Table 6. The Cepheid distance (with CTE correction) to nearby galaxies with different Galactic PL relations. GFG98a Galaxy 29.765 ± 0.061 NGC 925 NGC 1326A 31.093 ± 0.090 NGC 1365 31.336 ± 0.053 NGC 2090 30.415 ± 0.038 NGC 2541 30.464 ± 0.075 27.867 ± 0.098 NGC 3031 NGC 3198 30.842 ± 0.061 NGC 3319 30.569 ± 0.096 NGC 3351 29.936 ± 0.090 NGC 3621 29.249 ± 0.060 30.929 ± 0.067 NGC 4321 NGC 4414 31.265 ± 0.063 NGC 4535 31.020 ± 0.055 30.846 ± 0.078 NGC 4548 NGC 4725 30.541 ± 0.067 NGC 7331 30.875 ± 0.091 28.285 ± 0.129 IC 4182 NGC 3627 30.237 ± 0.083 31.726 ± 0.139 NGC 3982 NGC 4496A 30.974 ± 0.043 NGC 4527 30.771 ± 0.132 NGC 4536 30.989 ± 0.102 31.812 ± 0.112 NGC 4639 NGC 5253 27.776 ± 0.243 29.377 ± 0.062 NGC 4258 FSG03a T03a 29.768 ± 0.060 31.070 ± 0.089 31.283 ± 0.052 30.392 ± 0.036 30.430 ± 0.072 27.854 ± 0.095 30.805 ± 0.060 30.560 ± 0.094 29.946 ± 0.090 29.226 ± 0.059 30.869 ± 0.067 31.204 ± 0.052 30.969 ± 0.054 30.834 ± 0.078 30.490 ± 0.066 30.858 ± 0.083 28.330 ± 0.128 30.206 ± 0.083 31.682 ± 0.136 30.926 ± 0.043 30.723 ± 0.129 30.929 ± 0.102 31.758 ± 0.109 27.846 ± 0.243 29.401 ± 0.058 29.833 ± 0.061 31.166 ± 0.091 31.414 ± 0.053 30.488 ± 0.039 30.539 ± 0.075 27.939 ± 0.098 30.917 ± 0.061 30.639 ± 0.096 30.004 ± 0.090 29.322 ± 0.061 31.008 ± 0.067 31.344 ± 0.064 31.097 ± 0.055 30.917 ± 0.078 30.618 ± 0.067 30.947 ± 0.092 28.346 ± 0.129 30.311 ± 0.083 31.802 ± 0.139 31.051 ± 0.044 30.848 ± 0.132 31.068 ± 0.102 31.890 ± 0.113 27.834 ± 0.242 29.442 ± 0.062 a GFG98 = Gieren et al. (1998) PL relations; FSG03 = Fouqu´e et al. (2003) PL relations; T03 = Tammann et al. (2003) PL relations. See Section 3.2.2 for details. 0.004mag., compared to the mean light counterparts), be- cause Cepheids will appear closer at maximum light. The errors associated with maximum light are generally com- parable to those at mean light. This will be studied in future work. 4.3. Comparison to Published Results In order to compare our results with published values, we need to select the representative distance moduli from both LMC or Galactic PL relations. We pick the distance moduli from the derived OGLE LMC PL relations in Sec- tion 3.2.1 (i.e. column 4 of Table 5) and FSG03 PL rela- tions to represent the LMC and the Galactic results, re- spectively. The CTE-corrected distance moduli, µ0,CT E, are listed in Table 8. However, the metallicity corrections are not required in these comparisons because they are the same for published results and this work. The effect of metallicity corrections are discussed in Section 4.4. The results of the comparisons are plotted in Figure 8 and the average difference between our distance mod- uli and the published values are presented in Table 11 for both the LMC and Galactic PL relations. The overall com- parisons indicate that our results are consistent with the published results. When using the LMC PL relation, our distance moduli agree well with the KP results, but show a large discrepancy for the STS results (−0.178mag.). A careful comparison of Table 5 with published results shows that this difference arises from two sources: ∼ 0.10mag. is due to the use of MF91 LMC PL relations by STS and ∼ 0.07mag. is due to differences in the V and I band means when calculating the distance moduli (see Section 4.1 and Table 10). However, when using the Galactic PL relations, our results are different by ∼ 0.1mag. compared to both KP and STS results. Clearly, the use of Galactic PL relations increases the distance moduli. Surprisingly, the distance modulus of NGC 4258 calculated from LMC PL relations is closer to the water maser distance although the metallicity of this galaxy is nearly identical to Solar. Gibson et al. (2000, hereafter G00) used KP techniques to reanalyze the STS galaxies. Since G00 did the photom- etry from scratch, their list of Cepheids in each target galaxy is in general different from those found by STS. Moreover, they used the Madore & Freedman (1991) cal- ibrating PL relations (MF91). Gibson & Stetson (2001) and Stetson & Gibson (2001) used the same photometry as G00 but used the Udalski et al. (1999a) LMC PL relations (U99) to re-calibrate the distance to all the STS galaxies. They found shorter distance than those published by STS. They suggested that the discrepancy between their work 14 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 32 31 30 29 28 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 28 29 30 31 32 28 29 30 31 32 Fig. 6. Upper panel: Comparison of the distance moduli (after cor- rected for CTE) obtained by us- ing the Fourier techniques to es- timate the mean magnitudes with the published values. The line rep- resents the case of µP ub,CT E = µHere,CT E. Lower panel: The plot of the difference, ∆ = µHere,CT E − µP ub,CT E, vs. the distance moduli. The dashed line is the average dif- ference of −0.02. The symbols are: crosses = 16 KP galaxies; open cir- cles = 8 STS galaxies; and filled cir- cle = WM galaxy. and that of STS is due to the use of U99 LMC PL re- lations, their photometric analysis of the raw data and the STS analysis to account for possible correlated mea- surement errors in V and I. While these are important issues, our approach is to use exactly the same photome- try as used by both KP and STS groups. Our methods are closer to KP, though we have a different way of estimating the V and I band means. Our results show that, even with exactly the same Cepheids as used by STS, our distance moduli are significantly shorter than STS but significantly longer than those published in Gibson & Stetson (2001). It would be interesting to perform our analysis on the G00 photometry but this is not available as advertised on the KP web-site. 4.4. Comparison of Metallicity-Corrected Distance Moduli As well as comparing our results with the published val- ues, we discuss the distance moduli obtained when LMC and Galactic PL relations are used. First, comparing the distance moduli that are uncorrected for metallicity, µ0,CT E(LM C) and µ0,CT E(GAL) in Table 8, show that the distance moduli from LMC PL relations are always shorter than their Galactic counterparts (see Figure 9(a)). Summaries of this comparison are given in Table 12. This shows that the LMC PL relations will give a smaller dis- tance modulus by ∼ 0.07± 0.01mag on average (or about 3.5% in distance) compared to the distance modulus ob- tained from Galactic PL relations. The negligible random errors suggest that the difference of 0.07mag. might be significant, although it is small. However, when the metal- licity corrected distance moduli from the two sets of PL relations (µz(LM C) and µz(GAL) in Table 8) are com- pared, the difference between them falls close to zero. The results are listed in Table 12 and shown in Figure 9(b). Another approach is to use the Galaxy and LMC (P> 10d) as calibrating PL relations for metal rich and metal poor galaxies respectively. Thus if the seven SN calibrators of STS were forced on an LMC PL relation (U99), one would obtain an average distance decrease of 0.17mag. However, if we force the five galaxies in their sample, which have on average the same metallicity as the Galaxy, onto the steep relation given in T03, the pub- lished STS distances are recovered for these galaxies to within 0.01mag. Applying the same procedure to the 10 KP galaxies which are metal rich increases their distance moduli by about 0.16mag. on average, whilst using TR02 for the remaining metal poor galaxies increases their dis- tance moduli by 0.02mag. It has been pointed out by an anonymous referee that it may not be appropriate to eliminate those Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 15 0.3 0.2 0.1 0 0.3 0.2 0.1 0 0.5 0.5 1 log(P) 1.5 1 log(P) 1.5 0.3 0.2 0.1 0 0.5 0.3 0.2 0.1 0 0.5 2 2 1 log(P) 1 log(P) 1.5 1.5 Fig. 7. Comparison of A2 from Fourier techniques and template methods. The upper panels show the values of A2 from direct Fourier fit with the method described in Ngeow et al. (2003), and the lower panels show the values of A2 by using the template light curves given by Stetson (1996). Since the template light curves are undefined for log(p) < 0.85, Cepheids with period shorter and longer than log(p) = 0.85 are represents as open circles and filled triangles, respectively. Left:(a) The "calibrating set" Cepheids. Right:(b) The OGLE LMC Cepheids. Table 7. The Cepheid distance to nearby galaxies with PL(Max) relations. Galaxy NGC 925 NGC 1326A NGC 1365 NGC 2090 NGC 2541 NGC 3031 NGC 3198 NGC 3319 NGC 3351 NGC 3621 NGC 4321 NGC 4414 NGC 4535 NGC 4548 NGC 4725 NGC 7331 IC 4182 NGC 3627 NGC 3982 NGC 4496A NGC 4527 NGC 4536 NGC 4639 NGC 5253 NGC 4258 OGLE LMC PL(Max)a 29.699 ± 0.059 30.994 ± 0.089 31.165 ± 0.051 30.309 ± 0.035 30.343 ± 0.071 27.793 ± 0.094 30.712 ± 0.062 30.479 ± 0.092 29.871 ± 0.087 29.147 ± 0.059 30.740 ± 0.066 31.099 ± 0.042 30.856 ± 0.053 30.750 ± 0.081 30.388 ± 0.066 30.787 ± 0.082 28.253 ± 0.130 30.116 ± 0.087 31.592 ± 0.143 30.840 ± 0.044 30.616 ± 0.128 30.843 ± 0.102 31.659 ± 0.112 27.773 ± 0.226 29.352 ± 0.057 a Corrected for the CTE effects of δCT E (taken from last columns of Table 4). stars with a low pulsational amplitudes shown in Fig- ure 3 as this may bias the LMC PL relation. Such low amplitude Cepheids do exist and are seen in ex- ternal galaxies. If we include this group of stars in deriving the LMC PL relations, the new PL relations are: MV = −2.713(±0.044) log(P ) − 1.418(±0.030) and MI = −2.943(±0.029) log(P ) − 1.899(±0.020). The dis- tance moduli derived from these new LMC PL relations is ∼ 0.01mag. shorter when compared to the distance mod- uli derived from LMC PL relations given in Section 3.2.1. When comparing the distance moduli from these new LMC PL relations to those derived using Galactic PL rela- tions, the average difference for the 25 target galaxies con- sidered here is −0.052±0.006mag. and −0.082±0.006mag. with and without the metallicity corrections respectively. These low amplitude stars are not first overtone pul- sators since their light curve Fourier parameters fall in the region occupied by fundamentals. They lie in the in- stability strip and do not have unusual colors for their period. Thus their luminosities, masses and temperatures are similar to other Cepheids of similar period. So why do they have such a low amplitude? It could be that they have a slightly different composition or are just entering or leaving the fundamental mode instability strip (Buch- ler & Koll´ath 2002). In this case it is our contention that they should be excluded from the sample since they are Cepheids undergoing a transition. The use of Galactic PL relations to calibrate the Cepheid distance scale has been tried before by Paturel et al. (2002a) and Paturel et al. (2002b). The first paper, Paturel et al. (2002a), used the GFG98 sample and the 16 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations method of "sosie" (Paturel 1984) to determine Cepheid distances without assuming the PL relations. The sec- ond paper, Paturel et al. (2002b), applied the Hippar- cos calibrated PL relations (Feast & Catchpole 1997) to find Cepheid distances. The mean offsets of their results and the published KP results are 0.161 ± 0.029mag. and 0.027 ± 0.016mag. for their first and second papers, re- spectively. The latter result is comparable to our work. However, the extra-galactic Cepheids used in both papers are slightly different than either the KP or STS groups, in contrast with our study in this paper, because these au- thors applied a different method to deal with incomplete- ness bias. Also, these authors did not apply any metallicity corrections in their papers. 5. Conclusion and Discussion In this paper we recalibrate the Cepheid distance to about two dozen nearby galaxies, including the KP and STS galaxies, and a water maser galaxy. We use much of the same methodology as the KP team: the same LMC distance, value of R and metallicity correction, and the same (number of) Cepheids along with the published pho- tometries and periods. However our approach is differ- ent from the KP team in two aspects: (a) we estimate the mean magnitudes of sparsely sampled HST data from Fourier techniques (Ngeow et al. 2003) by reconstructing the Cepheid light curves; and (b) we use different sets of PL relations, including the new OGLE LMC PL relations, the Galactic PL relations and the LMC PL(Max) rela- tions, in distance determination. Overall, our results are consistent with each other and KP. We find significantly shorter distances to the STS galaxies. The use of the new Fourier techniques to obtain V and I band means does not produce significant deviations from the existing methods. However the derived distance modulus is more sensitive to the calibrating PL relation that is adopted. The most striking result from this study is that the distance moduli derived from using the LMC and Galactic PL relations are indistinguishable after the metallicity corrections are ap- plied. This provides strong support for the size and quan- tity of the metallicity dependence of the Cepheid PL rela- tions and the non-universality of PL relations (Tammann et al. 2003). Recent work has shown that the PL relation in the Galaxy and LMC have different slopes and moreover that the PL relation in the LMC is "broken" at a period of 10 days. Metallicity affects the mean brightness of a Cepheid and up to now, this has been the standard motivation for deriving and applying simple additive corrections to the PL relation to account for metallicity differences be- tween calibrating and target galaxy. However it may be that metallicity also affects the slope of the PL relation. If this is so, then the use of a simple additive correction as given in equation (2) is not really appropriate because the PL relation may have a different slope in the target and calibrating galaxy. However, the right panels of figure 8 show clearly that it works in the sense that the distance moduli differences obtained when using the Galactic and LMC as calibrating PL relations are driven close to zero. We try to investigate why this occurs in what follows. If we want to compare the distance modulus obtained using LMC and Galactic calibrating PL relations (or any pair of PL relations) when using exactly the same (number of) Cepheids in the target galaxy, including the same peri- ods and mean magnitudes, then the difference in distance modulus can be expressed as: ∆µ0 = [(R − 1)∆aV − R∆aI ]P log(P )i +(R − 1)∆bV − R∆bI, N (5) where ∆a(V,I) and ∆b(V,I) are the differences in slopes and zero-points for the two PL relations, respectively. Then, for these two PL relations, the change in the dis- tance modulus is a simple linear function of the target galaxy period distribution (< log(P ) >≡ P log(P )i/N , hereafter mean period), under the assumption of constant R (i.e. the universality of Galactic extinction law, see ob- servational verification for this assumption by Macri et al. 2001). The error in ∆µ0 is the quadrature sum of the σµ from both PL relations for the target galaxy. If metallicity corrections of δz (from equation (2)) are ap- plied to equation (5), the metallicity of the target galaxies cancels out and leaves the difference between the LMC and Galactic metallicity. Adopting [O/H]LM C = 8.50dex and [O/H]GAL = 8.87dex, along with γ = −0.2 ± 0.2mag. dex−1, the difference in metallicity-corrected dis- tance modulus becomes: ∆µz = ∆µ0 − 0.2(8.50 − 8.87) = ∆µ0 + 0.074 (±0.074) (6) It is worth pointing out that equations (5) and (6), though straightforward to derive, have not been presented in the literature before to the best of our knowledge. Equation (6) is of course dependent on the metallicity correction law adopted and the slopes of the calibrat- ing PL relations. We can re-write equation (5) as ∆µ0 = c < log(P ) > +d, where ∆µ0 = µ0(LM C) − µ0(GAL), c = (R− 1)∆aV − R∆aI and d = (R− 1)∆bV − R∆bI. For the four LMC PL relations (excluding the PL(Max) re- lations) and the three Galactic PL relations considered in this paper, there are a total of 12 combinations of (LM C, GAL) PL relations. The coefficients of c and d for each combination are listed in column 2 & 3 of Table 13. In the same table, we also list out the coefficients of d after applying the metallicity corrections (i.e. equation (6)), i.e. ∆µz = c < log(P ) > +dz where dz = d + 0.074. To see what values of < log(P ) > would produce identical distance moduli from using either the LMC or Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 17 Table 8. Comparisons of the distance moduli to target galaxies after CTE and metallicity corrections. µ0,CT E (LM C) Galaxy 29.718 ± 0.060 NGC 925 NGC 1326A 30.998 ± 0.088 NGC 1365 31.186 ± 0.051 NGC 2090 30.319 ± 0.036 NGC 2541 30.348 ± 0.071 27.791 ± 0.093 NGC 3031 NGC 3198 30.721 ± 0.060 NGC 3319 30.501 ± 0.092 NGC 3351 29.902 ± 0.090 29.155 ± 0.059 NGC 3621 NGC 4321 30.764 ± 0.067 NGC 4414 31.099 ± 0.044 NGC 4535 30.874 ± 0.053 30.773 ± 0.078 NGC 4548 NGC 4725 30.393 ± 0.067 NGC 7331 30.791 ± 0.079 28.318 ± 0.130 IC 4182 NGC 3627 30.127 ± 0.084 NGC 3982 31.593 ± 0.134 NGC 4496A 30.833 ± 0.043 30.629 ± 0.128 NGC 4527 NGC 4536 30.826 ± 0.102 NGC 4639 31.659 ± 0.106 27.854 ± 0.246 NGC 5253 29.370 ± 0.056 NGC 4258 µz(LM C) 29.728 30.998 31.287 30.379 30.348 27.841 30.741 30.477 30.050 29.205 30.890 31.239 31.014 30.941 30.477 30.825 28.298 30.277 31.673 30.887 30.709 30.896 31.759 27.784 29.440 µ0,CT E (GAL) 29.768 ± 0.060 31.070 ± 0.089 31.283 ± 0.052 30.392 ± 0.036 30.430 ± 0.072 27.854 ± 0.095 30.805 ± 0.060 30.560 ± 0.094 29.946 ± 0.090 29.226 ± 0.059 30.869 ± 0.067 31.204 ± 0.052 30.969 ± 0.054 30.834 ± 0.078 30.490 ± 0.066 30.858 ± 0.083 28.330 ± 0.128 30.206 ± 0.083 31.682 ± 0.136 30.926 ± 0.043 30.723 ± 0.129 30.929 ± 0.102 31.758 ± 0.243 27.846 ± 0.243 29.401 ± 0.058 µz(GAL) µ0,CT E (P U B)a µz(P U B)a 29.704 30.996 31.301 30.378 30.356 27.830 30.751 30.462 30.020 29.202 30.921 31.270 31.035 30.928 30.500 30.818 28.236 30.282 31.688 30.906 30.783 30.925 31.784 27.702 29.397 29.80 ± 0.04 31.04 ± 0.10 31.18 ± 0.05 30.29 ± 0.04 30.25 ± 0.05 27.75 ± 0.08 30.68 ± 0.08 30.64 ± 0.09 29.85 ± 0.09 29.08 ± 0.06 30.78 ± 0.07 31.10 ± 0.05 30.85 ± 0.05 30.88 ± 0.05 30.38 ± 0.06 30.81 ± 0.09 28.36 ± 0.09b 30.22 ± 0.12 31.72 ± 0.14 31.03 ± 0.14 30.72 ± 0.12 31.10 ± 0.13 32.03 ± 0.22 28.08 ± 0.20c 29.40 ± 0.09 29.81 31.04 31.27 30.35 30.25 27.82 30.70 30.62 30.00 29.11 30.91 31.24 30.99 31.05 30.46 30.84 · · · · · · · · · · · · · · · · · · · · · · · · 29.47 a Published distance moduli for KP galaxies are taken from Freedman et al. (2001), Table 4. The distance moduli for STS galaxies and WM galaxy are taken from series of STS papers and Newman et al. (2001), respectively. b Saha et al. (1994) assumed AV = AI = 0, hence µ = (µV + µI )/2. c The paper (Saha et al. 1995) did not list out the final µ, hence this value is calculated via equation (2) with µV and µI given in the paper. 32 31 30 29 28 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 32 31 30 29 28 28 29 30 31 32 28 29 30 31 32 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 28 29 30 31 32 28 29 30 31 32 Fig. 8. Comparisons with published results. The symbols are same as in Figure 6. Left:(a) Comparison of our results from the new OGLE LMC PL relations. The dashed line is the average difference of −0.06. Right:(b) Comparison of our results from the Galactic PL relations. The dashed line is the average difference of 0.01. 18 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 32 31 30 29 28 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 32 31 30 29 28 28 29 30 31 32 28 29 30 31 32 0.6 0.4 0.2 0 -0.2 -0.4 -0.6 28 29 30 31 32 28 29 30 31 32 Fig. 9. Comparisons of distance moduli from LMC and Galactic PL relations. The symbols are same as in Figure 6. Left:(a) Uncorrected for metallicity. The dashed line is the average difference of −0.07. Right:(b) Corrected for metallicity. The dashed line is the average difference of −0.00. Table 9. Comparisons of the mean magnitudes with pub- lished results. Table 10. Comparisons of the distance moduli with dif- ferent mean magnitudesa. ∆V /N a ∆I/N a Galaxy NGC 925 NGC 1326A 15 NGC 1365 NGC 2090 NGC 2541 NGC 3031 NGC 3198 NGC 3319 NGC 3351 NGC 3621 NGC 4321 NGC 4414 NGC 4535 NGC 4548 NGC 4725 NGC 7331 IC 4182 NGC 3627 NGC 3982 NGC 4496A 45 NGC 4527 NGC 4536 NGC 4639 NGC 5253 NGC 4258 0.079 ± 0.004 0.113 ± 0.015 0.065 ± 0.022 0.001 ± 0.059 0.046 ± 0.006 0.035 ± 0.029 0.062 ± 0.029 0.018 ± 0.009 0.027 ± 0.007 0.066 ± 0.008 0.087 ± 0.021 0.024 ± 0.011 0.086 ± 0.010 0.050 ± 0.013 0.045 ± 0.008 N 72 −0.046 ± 0.004 −0.006 ± 0.010 0.007 ± 0.006 47 −0.077 ± 0.006 −0.029 ± 0.014 0.064 ± 0.007 30 0.086 ± 0.020 29 0.039 ± 0.014 16 36 0.020 ± 0.063 −0.001 ± 0.012 33 0.009 ± 0.012 48 0.072 ± 0.008 59 −0.005 ± 0.013 42 8 0.077 ± 0.081 0.051 ± 0.009 47 0.023 ± 0.012 24 0.023 ± 0.008 15 13 0.077 ± 0.031 −0.011 ± 0.020 26 25 −0.018 ± 0.026 14 −0.018 ± 0.014 −0.030 ± 0.014 −0.016 ± 0.019 13 −0.013 ± 0.016 −0.010 ± 0.020 −0.005 ± 0.015 31 −0.007 ± 0.017 15 −0.071 ± 0.045 5 15 0.007 ± 0.012 0.040 ± 0.009 0.043 ± 0.004 0.057 ± 0.005 0.041 ± 0.009 0.026 ± 0.007 0.049 ± 0.018 Case 16 KP + WM galaxies 8 STS galaxies All 25 galaxies ∆(mag.)b 0.005 ± 0.015 −0.075 ± 0.037 −0.020 ± 0.017 a By using the same PL relations but different mean mag- nitudes. See Section 4.1 for details. b Mean Difference, ∆ =< µ0,CT E (Here)−µ0,CT E(P ub.) >, the errors are the standard deviations of the means. Table 11. Comparisons with the published results. Case ∆(mag.)a 16 KP galaxies 8 STS galaxies All 25 galaxies 16 KP galaxies 8 STS galaxies All 25 galaxies Using LMC PL relations −0.002 ± 0.017 −0.178 ± 0.039b −0.059 ± 0.023 Using Galactic PL relations 0.075 ± 0.018 −0.108 ± 0.038 0.014 ± 0.024 a Mean Difference, ∆ =< µ0,CT E (Here)−µ0,CT E(P ub.) >, the errors are the standard deviations of the means. b This difference is mainly due to the different LMC PL relations used. See text (Section 4.3) for details. b ∆V = VHere − VP ublished. Same for ∆I. the Galactic PL relations, we solve for < log(P ) > such that ∆µ0 = 0 or ∆µz = 0. The solutions are listed in columns 5 & 6 in Table 13 for the case of ∆µ0 = 0 and ∆µz = 0, respectively. From Table 13, the mean pe- riod for two distance moduli to be identical is around Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 19 Table 12. Comparisons of metallicity-corrected distance moduli. Case ∆(mag.)a 16 KP galaxies 8 STS galaxies All 25 galaxies 16 KP galaxies 8 STS galaxies All 25 galaxies Without metallicity correction −0.077 ± 0.005 −0.070 ± 0.015 −0.073 ± 0.006 With metallicity correction −0.002 ± 0.005 −0.003 ± 0.018 −0.001 ± 0.007 a Mean Difference, ∆ =< µ(LM C) −µ(GAL) >, the errors are the standard deviations of the means. < log(P ) >0∼ 1.1 − 1.2, and < log(P ) >z∼ 1.3 − 1.4, except for the (MF91,GFG98) and (MF91,FSG03) pairs. In Table 14, we list out the observed mean periods for the target galaxies. This shows that most of the extra- galactic Cepheids in our target galaxies have an observed mean period of ∼ 1.4 (the mean period for 16 KP galaxies is 1.418; the mean period for 8 STS galaxies is 1.401; and the mean period for all 25 galaxies is 1.342). Therefore, without applying the metallicity correction, the Galactic PL relations will be expected to produce longer distances than the LMC PL relations in most of the (LMC,GAL) pair of PL relations. For example, the change of the dis- tance modulus is ∼ 0.10mag. when comparing the LMC PL relations (either U99 or the PL relations derived in Section 3.2.1) to the FSG03 Galacic PL relations in in- dividual target galaxies. However, the LMC and Galactic PL relations will produce almost identical distance mod- uli after a metallicity correction to within ∼ 0.03mag., for most of the target galaxies. The only exception is NGC 5253, because the mean period for this galaxy is 1.029, which is much smaller than the required < log(P ) > of ∼ 1.4. The T03 Galactic PL relation has a different slope to the GFG03 and FSG03 Galactic PL relations. Thus the median period in Table 13 required for a simple additive correction to be sufficient is approximately 1.2. This is slightly outside the range of mean period required to pro- duce similar distance moduli after a simple metallicity correction when both Galactic and LMC calibrating PL relations are used. If the result of this paper, i.e. the near identical dis- tance moduli from LMC and Galactic PL relations af- ter metallicity corrections, is true, then this result can be used to constrain the sign for the coefficient of metal- licity correction, the γ in equation (2). The value of γ = 0.2±0.2mag dex−1 used in this study is adopted from Freedman et al. (2001), which is roughly the midrange value from several empirical studies (see Section 3). Since the difference of distance moduli between LMC and Galac- tic PL relations is about −0.07mag. without the metallic- ity correction (Table 12), a +0.07mag. metallicity correc- tion is required to bring the two distance moduli to be identical. The correction of 0.074mag. in equation (6) is almost identical to this requirement, hence the value of γ should be around −0.2mag dex−1 and constrains the sign to be negative. In addition, if the Cepheid PL relations do indeed depend on metallicity, the result of this paper suggests that a simple additive metallicity correction as in equation (2) is a good approximation to model the full complexity of the metallicity dependence of the Cepheid PL relation, provided the mean period of Cepheids in the target galaxy are in the appropriate range for the slopes and extinction laws adopted for the calibrating PL relations6. Some researchers suggest using the LMC PL relations and Galactic PL relations for metal-poor and metal-rich galaxies, respectively, forgetting about metal- licity corrections until an solid understanding of this topic is obtained. This naturally begs the question of which cal- ibrating PL relation to use if the metallicity of the target galaxy is in between the LMC and Galaxy. However our result does not, at the moment, provide evidence support- ing one Galactic PL relation over another. In summary, the above discussion suggests that: (a) metallicity corrections are necessary when using Cepheid PL relations to find distance moduli; (b) as a consequence, the PL relations do depend on metallicity; (c) hence, Cepheids in the LMC and Galaxy obey different PL rela- tions; (d) the sign for the coefficient of metallicity correc- tion (γ) has to be negative; and (e) both LMC and Galac- tic PL relations can be used to determine the distance modulus because either one of the PL relations would yield the same distance modulus after the appropriate metallic- ity correction. However, further study is needed to test these conclusions. Acknowledgements. This work has been supported by NASA Grant GO-09155.03. Part of SN work was performed under the auspices of the U.S. Department of Energy, National Nu- clear Security Administration by the University of Califor- nia, Lawrence Livermore National Laboratory under contract W7405-Eng-48. We thank an anonymous referee for many help- ful suggestions that made the paper more relevant. Special thanks to D. Leonard for some discussions and G. A. Tam- mann for letting us to read their unpublished papers and useful discussion. The authors would like to thank R. Ciardullo for providing the reference of Solar metallicity. References Beaulieu, J. et al., 1997, A&A, 318, L47 Beaulieu, J. P., Buchler, J. R. & Koll´ath, Z., 2001, A&A, 373, 164 Buchler, R. & Koll´ath, Z., 2002, ApJ, 573, 324 Caputo, F., Marconi, M., Musella, I. & Santolamazza, P., 2000, A&A, 359, 1059 Cardelli, J. A., Clayton. G. C. & Mathis, J. S., 1989, ApJ, 345, 245 6 We thank the referee to point out this. 20 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations Table 13. The change of ∆µ with different pairs of LMC and Galactic PL relationsa. (LMC-GAL) pair (MF91-GFG98) (U99-GFG98) (Here-GFG98) (TR02-GFG98) (MF91-FSG03) (U99-FSG03) (Here-FSG03) (TR02-FSG03) (MF91-T03) (U99-T03) (Here-T03) (TR02-T03) c d dz = d + 0.074 < log(P ) >0 < log(P ) >z -0.257 -0.497 -0.470 -0.439 -0.006 -0.246 -0.218 -0.188 -0.300 -0.540 -0.513 -0.482 0.369 0.608 0.561 0.548 0.041 0.280 0.233 0.220 0.365 0.595 0.548 0.534 0.443 0.682 0.635 0.622 0.115 0.354 0.307 0.294 0.430 0.669 0.622 0.608 1.4343 1.2231 1.1946 1.2465 6.8417b 1.1392 1.0673 1.1678 1.1859 1.1018 1.0692 1.1085 1.7218 1.3719 1.3521 1.4149 19.175b 1.4399 1.4060 1.5614 1.4325 1.2388 1.2135 1.2620 a ∆µ = c < log(P ) > +d (i.e. equation (5)), and < log(P ) > is the mean period such that ∆µ = 0. b The large values of < log(P ) > are due to the small value of c in column 2. Hence the MF91 LMC PL relations will always produce a larger distance modulus, as compared to distance modulus from FSG03 Galactic PL relations. Table 14. The mean period in target galaxies. Galaxy NGC 925 NGC 1326A NGC 1365 NGC 2090 NGC 2541 NGC 3031 NGC 3198 NGC 3319 NGC 3351 NGC 3621 NGC 4321 NGC 4414 NGC 4535 NGC 4548 NGC 4725 NGC 7331 IC 4182 A NGC 3627 NGC 3982 NGC 4496A NGC 4527 NGC 4536 NGC 4639 NGC 5253 NGC 4258 < log(P ) > 1.2937 1.3962 1.5152 1.3987 1.4423 1.3577 1.4525 1.3398 1.2681 1.3943 1.5454 1.5467 1.5050 1.3467 1.5081 1.3733 1.1244 1.4294 1.4774 1.4945 1.4971 1.5426 1.5201 1.0292 1.2084 Feast, M. & Walker, A., 1987, ARA&A, 25, 345 Feast, M. & Catchpole, R., 1997, MNRAS, 286, L1 Feast, M., 1999, PASP, 111, 775 Feast, M., 2001, Odessa Astronomical Pub., 14, 144, astro-ph/0110360 Feast, M., 2003, to be published in Stellar Candles, Lecture Notes in Physics, astro-ph/0301100 Freedman, W., 1988, ApJ, 326, 691 Freedman, W. & Madore, B., 1990, ApJ, 365, 186 Freedman, W. et al., 1994, ApJ, 427, 628 Ferrarese, L. et al., 2000, ApJS, 128, 431 Fouqu´e, P., Storm, J. & Gieren, W., 2003, to be pub- lished in Stellar Candles, Lecture Notes in Physics, astro-ph/0301291 Freedman, W. et al.,2001, ApJ, 533, 47 (KP) Gibson, B. et al., 2000, ApJ, 529, 723 Gibson, B. & Stetson, P., 2001, ApJ, 547, L103 Gieren, W., Fouqu´e, P. & G´omez, M., 1998, ApJ, 496, 17 Gould, A., 1994, ApJ, 426, 542 Grevesse, N., Noels, A. & Sauval, A. J., 1996, in Cosmic Abun- dances, ASP Conf. Series Vol. 99, eds. Holt & Sonneborn, pg 117 Groenewegen, M. & Oudmaijer, R. D., 2000, A&A, 356, 849 Groenewegen, M., 2000, A&A, 363, 901 Hendry, M. & Kanbur, S., 1996, in Mapping, Measuring, and Modeling the Universe, ASP Conf. Series Vol. 94, eds. Coles, Martinez & Pons-Borderia, pg. 357 (astro/ph- 9603014) Herrnstein, J. et al., 1999, Nature, 400, 539 Hill, R. et al., 1998, ApJ, 496, 648 Holtzman, J. A. et al., 1995, PASP, 107, 1065 Kanbur, S. & Hendry, M., 1996, A&A, 305, 1 Kelson, D. et al., 1996, ApJ, 463, 26 Kennicutt, R., Freedman, W. & Mould, J., 1995, AJ, 110, 1476 Kennicutt, R. et al., 1998, ApJ, 491, 13 Kochanek, C. S., 1997, ApJ, 491, 13 Labhardt, L., Sandage, A. & Tammann, G. A., 1997, A&A, 322, 751 Lanoix, P., Paturel, G. & Garnier, R., 1999a, ApJ, 517, 188 Lanoix, P., Paturel, G. & Garnier, R., 1999b, MNRAS, 308, 969 Leonard, D., Kanbur, S., Ngeow, C. & Tanvir, N., 2003, ApJ Submitted Macri, L. et al., 2001, ApJ, 549, 721 Madore, B., 1982, PASP, 253, 575 Madore, B. & Freedman, W., 1991, PASP, 103, 933 Moffett, T. & Barnes, T., 1984, ApJS, 55, 389 Moffett, T. & Barnes, T., 1986, ApJ, 304, 607 Mould, J. et al., 2000, ApJ, 529, 786 Newman, J. et al., 2001, ApJ, 553, 562 Kanbur et al.: Cepheid distance from LMC and Galactic PL relations 21 Ngeow, C., Kanbur, S., Nikolaev, S., Tanvir, N. & Hendry, M., 2003, ApJ, 586, 959 Nikolaev, S., Drake, A. J., Keller, S. C., Cook, K. H., Dalal, N., Griest, K., Welch, D. L. & Kanbur, S., 2003, ApJ Sub- mitted Paturel, G., 1984, ApJ, 282, 382 Paturel, G., Theureau, G., Fouqu´e, P., Terry, J. N., Musella, I. & Ekholm, T., 2002a, A&A, 383, 398 Paturel, G., Teerikorpi, P., Theureau, G., Fouqu´e P., Musella, I. & Terry, J. N., 2002a, A&A, 389, 19 Saha, A., Labhardt, L., Schwengeler, H., Macchetto, D., Pana- gia, N., Sandage, A. & Tammann, G. A., 1994, ApJ, 425, 14 Saha, A., Sandage, A., Labhardt, L., Schwengeler, H., Tam- mann, G. A., Panagia, N. & Macchetto, F. D., 1995, ApJ, 438, 8 Saha, A., Sandage, A., Labhardt, L., Tammann, G. A., Mac- chetto, F. D. & Panagia, N., 1996, ApJ, 466, 55 Saha, A., Sandage, A., Thim, F., Labhardt, L., Tammann, G. A., Christensen, J., Panagia, N. & Macchetto, F. D., 2001a, ApJ, 551, 973 Saha, A., Sandage, A., Tammann, G. A., Dolphin, A., Chris- tensen, J., Panagia, N. & Macchetto, F. D., 2001b, ApJ, 562, 314 (STS) Sandage, A. & Tammann, G. A., 1968, ApJ, 151, 531 Sandage, A., 1988, PASP, 100, 935 Sasselov, D. D. et al., 1997, A&A, 324, 471 Schechter, P., Mateo, M. & Saha, A., 1993, PASP, 105, 1324 Sebo, K, et al., 2002, ApJS, 142, 71 Silbermann, N. et al., 1996, ApJ, 470, 1 Simon, N., Kanbur, S. & Mihalas, D., 1993, ApJ, 414, 310 Spergel, D. N. et al., 2003, astro-ph/0302209, ApJ Submitted Stetson, P., 1994, PASP, 106, 250 Stetson, P., 1996, PASP, 108, 851 Stetson, P., 1998, PASP, 110, 1448 Stetson, P., et al., 1998b, ApJ, 508, 491 Stetson, P. & Gibson, B., 2001, MNRAS, 328, L1 Tammann, G. A., Reindl, B., Thim, F., Saha, A. & Sandage, A., 2001, in A New Era in Cosmology, ASP Conf. Series Vol. 283, eds. Metcalfe & Shanks, pg 258, astro-ph/0112489 Tammann, G. A. & Reindl, B., 2002, Ap&SS, 280, 165, astro-ph/0208178 Tammann, G. A., 2003, private-communication Tammann, G. A., Sandage, A. & Reindl, B., 2003, A&A in- press, astro-ph/0303378 Tanvir, N., 1997. in The Extra-galactic Distance Scale, STScI Series, eds. Livio, pg 91 (astro-ph/9611027) Tanvir, N., 1999, in Harmonizing Cosmic Distance Scales in a Post-HIPPARCOS Era, ASP Conf. Series Vol. 167, eds. Egret & Heck, pg 84 (astro-ph/9812356) Tanvir, N., Ferguson, H. & Shanks, T., 1999, MNRAS, 310, 175 Thim, F., Tammann, G. A., Saha, A., Dolphin, A., Sandage, A., Tolstoy, E. & Labhardt, L., 2003, ApJ Accepted, astro-ph/3030101 Udalski, A., Szymanski, M., Kubiak, M., Pietrzynski, G., Soszynski, I., Wozniak, P. & Zebrun, K, 1999a, AcA, 49, 201 Udalski, A., Soszynski, I., Szymanski, M., Kubiak, M., Pietrzynski, G., Wozniak, P. & Zebrun, K., 1999b, AcA, 49, 223 Willick, J. & Batra, P., 2000, ApJ, 548, 564
astro-ph/0607239
2
0607
2006-07-12T22:24:09
A Strong-Lens Survey in AEGIS: the influence of large scale structure
[ "astro-ph" ]
We report on the results of a visual search for galaxy-scale strong gravitational lenses over 650 arcmin^2 of HST/ACS imaging in the DEEP2-EGS field. In addition to a previously-known Einstein Cross (the "Cross," HST J141735+52264, with z_lens=0.8106 and a published z_source=3.40), we identify two new strong galaxy-galaxy lenses with multiple extended arcs. The first, HST J141820+52361 (the ``Dewdrop''; z_lens=0.5798, lenses two distinct extended sources into two pairs of arcs z_source=0.while), 9818 the second, HST J141833+52435 (the ``Anchor''; z_lens=0.4625), produces a single pair of arcs (source redshift not yet known). All three definite lenses are fit well by simple singular isothermal ellipsoid models including external shear. Using the three-dimensional line-of-sight (LOS) information on galaxies from the DEEP2 data, we calculate the convergence and shear contributions, assuming singular isothermal sphere halos truncated at 200 h^-1 kpc. These are also compared against three-dimensional local-density estimates. We find that even strong lenses in demonstrably underdense local environments may be considerably affected by LOS contributions, which in turn, may be underestimates of the effect of large scale structure.
astro-ph
astro-ph
Submitted to the AEGIS ApJL Special Issue Preprint typeset using LATEX style emulateapj v. 11/26/04 A STRONG-LENS SURVEY IN AEGIS: THE INFLUENCE OF LARGE SCALE STRUCTURE Leonidas A. Moustakas1, Phil Marshall2, Jeffrey A. Newman3 ,11, Alison L. Coil4 ,12, Michael C. Cooper5, Marc Davis5, Christopher D. Fassnacht6, Puragra Guhathakurta8, Andrew Hopkins9, Anton Koekemoer10, Nicholas P. Konidaris8, Jennifer M. Lotz11 ,13, and Christopher N. A. Willmer4 Submitted to the AEGIS ApJL Special Issue ABSTRACT We report on the results of a visual search for galaxy-scale strong gravitational lenses over 650 arcmin2 of HST /ACS imaging in the Extended Groth Strip (EGS). These deep F606W- and F814W-band observations are in the DEEP2-EGS field. In addition to a previously-known Einstein Cross also found by our search (the "Cross," HST J141735+52264, with zlens = 0.8106 and a published zsource = 3.40), we identify two new strong galaxy-galaxy lenses with multiple extended arcs. The first, HST J141820+52361 (the "Dewdrop"; zlens = 0.5798), lenses two distinct extended sources into two pairs of arcs (zsource = 0.9818 by nebular [O II] emission), while the second, HST J141833+52435 (the "Anchor"; zlens = 0.4625), produces a single pair of arcs (source redshift not yet known). Four less convincing arc/counter-arc and two-image lens candidates are also found and presented for complete- ness. All three definite lenses are fit reasonably well by simple singular isothermal ellipsoid models including external shear, giving χ2 ν values close to unity. Using the three-dimensional line-of-sight (LOS) information on galaxies from the DEEP2 data, we calculate the convergence and shear contri- butions κlos and γlos to each lens, assuming singular isothermal sphere halos truncated at 200 h−1 kpc. These are compared against a robust measure of local environment, δ3, a normalized density that uses the distance to the third nearest neighbor. We find that even strong lenses in demonstrably underdense local environments may be considerably affected by LOS contributions, which in turn, under the adopted assumptions, may be underestimates of the effect of large scale structure. Subject headings: gravitational lensing -- galaxies: high-redshift -- large-scale structure of uni- individual (HST J141735+52264) -- galaxies: verse -- galaxies: (HST J141820+52361) -- galaxies: individual (HST J141833+52435) individual 1. INTRODUCTION Galaxy-scale gravitational lenses have many astro- physical and cosmological applications. These rely on the ability to construct robust and accurate gravitational lens models. However, the contribution of the large-scale structure along the line of sight (LOS) between the observer and the source is often unknown, though it may be significant. In particular, though lens models may detect the influence of the distorting effects of environmental shear (γ) in a preferred direction, models of even the most richly-constrained Einstein Rings with Hubble Space Telescope images (e.g. Dye & Warren 2005; Wayth et al. 2005; Koopmans et al. 2006) are 1 JPL/Caltech, 4800 Oak Grove Dr, MS 169-327, Pasadena, CA 91109 [email protected] 2 KIPAC, P.O. Box 20450, MS29, Stanford, CA 94309 [email protected] 3 INPA, LBNL, Berkeley, CA [email protected] 4 S.O., University of Arizona, Tucson, AZ 85721 acoil, [email protected] 5 Department of Astronomy, U.C. Berkeley, Berkeley, CA 94720 cooper, [email protected] 6 Department of Physics, U.C. Davis, Davis, CA 95616 [email protected] 8 U.C.O./Lick Observatory, UCSC, Santa Cruz, CA 95064 raja, [email protected] 9 School of Physics, University of Sydney NSW, Australia [email protected] 10 Space Telescope Science Institute, Baltimore, MD 21218 [email protected] 11 NOAO, 950 North Cherry Street, Tucson, AZ 85719 [email protected] 12 Hubble Fellow 13 Goldberg Fellow the 2006; Indeed, still subject to the mass-sheet degeneracy due to extra field convergence (κ), which can lead to incorrect lens masses (e.g. Kochanek 2004). lens galaxies are often massive early-type galaxies, which are gen- erally found in groups or clusters. The most famous example is two-image lensed QSO Q0957+561 (Walsh et al. 1979; Young et al. 1980). The deter- mination of H0 from this system depends crucially on correctly modeling the galaxy cluster surrounding the primary lensing galaxy (e.g. Keeton et al. 2000). Several other lens-galaxy groups and environments have been studied in detail (Kundic et al. 1997a,b; Tonry 1999; Fassnacht & Lubin 1998; Tonry & Kochanek 2005; 2002; Morgan et al. Momcheva et al. 2006a; Auger et al. 2006), with sometimes inconclusive re- sults. In analyses such as in Keeton & Zabludoff (2004), through mock lens realizations, it is shown how local environment may affect key applications of lenses. They argue that H0 and ΩΛ may be overestimated, the expected ratio of four-image to two-image lenses may be underestimated, and predictions for millilensing by dark matter substructure may be off by significant amounts. Other theoretical work (Bar-Kana 1996; Metcalf 2005; Wambsganss et al. 2005) suggests that all matter along a line of sight can be important. 2005; Williams et al. Fassnacht et al. In the emergent era of large-solid angle, densely- sampled spectroscopic surveys that may include strong lenses, both environmental and large scale structure ef- fects can be explored quantitatively. The DEIMOS spec- troscopy of the Extended Groth Strip (EGS) is particu- 2 Moustakas et al. larly well-suited to this task, and is employed here to both discover new strong galaxy-lenses, and to begin ad- dressing the quantitative effect of environment in their behavior. The DEEP2-EGS field is a 120×30 arcmin strip, the focus of the "All-wavelength EGS International Survey" (AEGIS), includes deep CFHT BRI imaging (Coil et al. 2004a) and Keck/DEIMOS spectroscopy of nearly 14 000 galaxies to date. The spectroscopy is ∼ 75% complete to RAB < 24.1. For the analysis here, we only employ the most certain redshift assignments (Coil et al. 2004b). Deep HST /ACS imaging of nearly 650 arcmin2 over 63 stitched tiles reach V606 = 28.75 and I814 = 28.10 (AB, 5σ point source; Davis et al, this issue). These data lend themselves to two different techniques for searching for heretofore-unknown gravitational lenses: spectroscopi- cally and visually. The spectroscopic redshifts are sup- plemented as necessary with photometric redshifts mea- sured from deep KPNO U BV RI imaging (A. Hopkins et al., in prep). The spectroscopic approach of searching for "anoma- lous" emission lines in early-type spectra has some his- tory (e.g. Warren et al. 1996), and has recently proved to be spectacularly successful when applied to SDSS spectroscopy (Bolton et al. 2004; Willis et al. 2005) with HST /ACS followup (Bolton et al. 2005, 2006; Treu et al. 2006). Explicitly spectroscopic searches for lenses in the DEEP2 data will be explored elsewhere. In the imaging domain, one may hope to search for lens candidates by some automated algorithm, or by vi- sual inspection (e.g. Ratnatunga et al. 1995; Zepf et al. 1997; Fassnacht et al. 2004). The more quantitative and objective automated approach may eventually be pre- ferred (especially for datasets larger than the one con- sidered here), but would, however, require a training set. The EGS ACS data described here is used for just this purpose in a separate work (Marshall et al. in prep) as a precursor to searching the entire HST imaging dataset.13 Towards that goal we have undertaken a search for lenses by purely visual inspection. The lens-search methodology is described in § 2. The newly discovered lenses and the modeling results are given in § 3, while measurements of the local and LOS environments of the lenses are given in § 4. Discus- sion and conclusions are the subject of § 5. A concor- dance flat cosmology with ΩΛ = 1 − Ωm = 0.7 and H0 = 100 h km s−1 Mpc−1 with h = 0.7 is used through- out. Unless otherwise stated, all magnitudes are in the AB system. 2. LENS-SEARCH METHODOLOGY The search for gravitational lens candidates was con- ducted by-eye. Three-color images of all of the ACS tiles were built following the Lupton et al. (2004) al- gorithm, using the photometric zeropoints to provide the relative scale factors, and using the mean of the F606W and F814W images for the green channel. The full ACS dataset was inspected repeatedly in the color images at full resolution, with plausible candidates clas- sified with grades of "A" or "B" and marked for fur- ther inspection. Object coordinates were then matched against the DEEP2 spectroscopic catalog, which includes 13 http://www.slac.stanford.edu/~pjm/HAGGLeS a "serendipitous feature" flag, for possible anomalous, higher-redshift emission lines. Emission from a source behind the Dewdrop lens (described below) was found in this way. Fig. 1. -- The three most plausible lenses from this survey, from left to right the Cross, the Dewdrop, and the Anchor (see text). From top to bottom, we show the discovery image, the lensed image model, the residuals by subtraction with the (lens-galaxy-removed) imaging data, and the reconstructed source. All panels, including the source-plane one, are approximately 3 arcsec on a side, with the exact dimensions shown by the scale bars. The image-plane critical curves and the source-plane caustics are shown in the third and fourth rows, respectively. Fig. 2. -- Additional lens candidates based on visual inspection. These are not yet bolstered by spectroscopy, but will be targetted when possible. The left two are candidate arc/counter-arc lenses, whereas the right two are candidate two-image lenses. Images are 3 arcsec square. 3. LENSES & MODELS In addition to a previously known Einstein Cross, we find two new unambiguous strong galaxy-galaxy lenses (Fig. 1). Four additional plausible lens candidates (Fig. 2) are also reported on. Here we describe the lens modeling and the model results for each lens. 3.1. Lens modeling and source reconstruction The lensed sources in the EGS all appear to be blue and extended, and are likely star forming galaxies at high Environments of gravitational lenses 3 redshift (z ∼ 1). We therefore take the image pixels as our data (rather than simply image-centroid positions), and predict the image using a simple ray tracing forward from the source plane, followed by a PSF convolution. We first subtract the lens galaxy light using a tilted 2D Moffat profile,14 and mask the very center of the lens galaxy where some residual flux remains. It is important that the unmasked region contain not only the lensed images but also the clean pixels that do not have lensed features. These clean pixels contain at least as much in- formation as the ones with lensed flux, vetoing models that predict images where there are none. For the pro- jected mass profile of the lens we adopt a singular isother- mal ellipsoid (SIE; Kormann et al. 1994) model, plus an external shear component. Using a Markov chain Monte Carlo procedure presented in detail elsewhere (Marshall et al. in prep), the position, ellipticity, orientation and mass of the lens, external shear amplitude and the direc- tion, position, ellipticity, orientation and Sersic profile parameters of the source are all fit to the data. Since we are interested in accurate estimation of the lens envi- ronment, we apply a prior on the orientation of the lens ellipticity to reflect the expected correlation with the lens light (e.g. Koopmans et al. 2006). 3.2. HST J141735+52264 (A1 -- Cross) lens was This originally discovered by Ratnatunga et al. (1995) by visual inspection of the HST /WFPC2 Medium Deep Survey (MDS) data. The lens redshift is zlens = 0.8106 (Table 1), and the source is at zsource = 3.4 (Crampton et al. 1996). The large Einstein radius θE = 1.447 arcsec and the four-image configuration require a large enclosed mass and a signif- icant amount of external shear, γmod = 0.080, a result consistent with Treu & Koopmans (2004). The best-fit model shows very small residuals at the two outer images, a feature corrected for by Treu & Koopmans (2004) with a potential gradient that is presumably associated with a nearby structure. The mass and external shear are not affected by this correction. 3.3. HST J141820+52361 (A2 -- Dewdrop) The Dewdrop lens at zlens = 0.5798 lenses two dis- tinct sources into two pairs of arcs. The Keck/DEIMOS spectrum of the system reveals anomalous [O II] nebu- lar emission at zsource = 0.9818 (Fig. 3). The sources in the Dewdrop system are part of a remarkable irregular and loose association of star formation knots and diffuse emitting material that extends over more than 10 arcsec, or more than 80 kpc comoving in size. 3.4. HST J141833+52435 (A3 -- Anchor) The Anchor system exhibits a pair of arcs created by a lens at a redshift of zlens = 0.4625. The best-fitting lens model requires a significant external shear contribution (see Table 1), as might be expected from the position and shape of the counter-image to the main arc. 14 The Moffat function is a modified Lorentzian with variable power law index. The fit is done with the MPFIT IDL suite of C. Markwardt. 4500 λ rest [A] 5000 5500 HSTJ141820.84+523611.2 (Dewdrop) z = 0.580 l 2000 1500 s t n u o c 1000 500 G band Hγ [O II] at zS=0.982 0 7000 7500 8000 λ obs [A] 8500 9000 Fig. 3. -- The DEIMOS spectrum of the Dewdrop lens clearly shows an "anomalous" doublet emission line (insert), which is read- ily identified as [O II] at zsource = 0.9818. 3.5. Additional lens candidates In Fig. 2 and Table 1 we identify four additional visually-identified lens candidates. Only two of the four presently have redshifts measured, and require further spectroscopic followup. These are presented for com- pleteness, and do not affect the scope or results of this paper. 4. THE ENVIRONMENTS OF THE LENSES We explore the environments of the lenses in two dif- ferent ways. The first makes use of a relatively unbi- ased measure of the very local environment of any one galaxy, dubbed δ3 and explored in detail in Cooper et al. (2005a,b). This parameter is derived from the distance to the third-nearest neighbor among the galaxies within 1000 km s−1 along the line of sight, and scales as the in- verse of the cube of this distance. Thus, more concen- trated environments have larger values of δ3. The typical uncertainties in individual measures of δ3 are ∼ 0.5 dex. We only compute this measure for galaxies with spectro- scopic redshifts. 1.000 0.100 0.010 ] γ [ r a e h s A2 A3 models los data A1 -3.0σ -2.0σ -1.0σ (U-B)0 > 0.9 B3 0.0σ B4 1.0σ 2.0σ 3.0σ r e b m u N 0.001 0.01 0.10 10.00 local density contrast [1 + δ 3] 1.00 100.00 Fig. 4. -- The lower panel shows the distribution of the local- enviromental measure 1+δ3 (Cooper et al. 2005a), for "red se- quence" galaxies with rest-frame colors (U − B)0 > 0.9. (All lenses are found to satisfy the same color-criterion). Based on a gaus- sian fit to the distribution, N σ positions are marked in the x-axis line above, as a guide. The upper panel shows the lens-model and line-of-sight shear for each object, as the filled-circles and open hexagons, respectively. The size of the hexagons corresponds to the calculated line of sight convergence κlos. The 1+δ3 values of the lens-candidates B3 and B4 are shown as well. 4 Moustakas et al. EGS lenses: data, environment, & models TABLE 1 Data Dec zlens R RA ID Alias (J2000) (J2000) M c B (AB) A1 Cross 14:17:35.72 52:26:46.3 0.8106 21.38 −21.25 A2 Dewdrop 14:18:20.77 52:36:11.3 0.5798 20.55 −20.35 A3 Anchor 14:18:33.11 52:43:52.6 0.4625 20.45 −19.47 B1 Flourish 14:18:07.32 52:30:29.8 0.847a 22.58 (−17.8) 14:20:52.01 53:06:57.2 0.601b 23.82 (−16.8) B2 Quotes 14:17:59.01 52:35:14.8 0.6863 21.50 −20.24 B3 Dots B4 Colon 14:20:53.89 53:06:07.0 0.3545 20.61 −18.47 (AB) zsource 3.40 0.9818 ... ... ... ... ... Environment log(1 + δ3) Nlos κlos γlos θγlos (◦E) +1.453 −1.260 −0.960 ... ... +0.109 +0.880 36 46 52 ... ... ... ... 0.17 0.02 78 0.10 0.02 140 0.09 0.02 146 ... ... ... ... ... ... ... ... ... ... ... ... σSIS θE (′′) km s−1 292.8 1.45 0.67 260.6 248.9 0.83 ... ... ... ... ... ... ... ... Models γmod θγmod (◦E) χ2 ν 0.080 115.3 1.081 0.071 101.4 1.005 0.153 140.8 0.933 ... ... ... ... ... ... ... ... ... ... ... ... a: σz = 0.067 & b: σz = 0.24 (A. Hopkins et al., in prep); c: parenthetical quantities are based on photometric redshifts. As a second probe of lens environment we model the contribution to the lensing potential due to individual neighboring galaxies using simple analytic mass distribu- tions. We calculate the convergence κlos and shear γlos line-of-sight contribution by all galaxies within a pro- jected separation of 200 h−1 kpc from the lens galaxies, out to the redshift of the source. We treat each galaxy as an isolated halo, undoubtedly neglecting the effect of group halos and other structures. Assuming that we can approximate each galaxy i as a singular isothermal sphere (SIS), we have κi = bi/2ri, where r is the projected dis- tance from the lens and b is the "lens strength" for a background source at angular diameter distances of Ds from the observer and Dls from the lens, b = 4π (cid:16) σdm c (cid:17)2 Dls Ds . (1) The central dark matter velocity dispersion σdm of each galaxy is assumed to be the same as the central stel- lar velocity dispersion, which is derived from the es- timated rest-frame B-band (Vega) magnitude of each galaxy using the Faber-Jackson relationship as given in Mitchell et al. (2005) (see also Jonsson et al. 2006). (We neglect the dispersion in this relation). The total shear contribution is the "headless-vector" sum of the shears, ¯γlos = Σ ¯γi, while the total convergence is a scalar sum: κlos = Σκi. It is worth noting that if at large radii the profiles are steeper than SIS (such as NFW), the conver- gence contribution will be smaller overall than the shear. These measurements are given in Table 1 and discussed in the last section. 5. DISCUSSION & CONCLUSIONS The numbers of definite lenses reported on here is con- sistent with other surveys. For example, Bolton et al. (2006) find that ∼0.1% of luminous red galaxies are very likely to be strong galaxy-galaxy lenses, although spe- cial lines of sight can have much higher lensing rates (e.g. Fassnacht et al. 2006b). The rate above then sug- gests that there should be ∼4 strong lenses in this survey, which is a good match to our three. The main conclusions of this work can be drawn by an examination of Fig. 4. The Cross is in a fairly overdense local environment, which is consistent with this lens being associated with the z ≈ 0.8 sheet de- scribed in Koo et al. (1996) and Im et al. (2002). Given this, the shear of ∼10% required by the model (see also Treu & Koopmans 2004) seems plausible. What seems surprising is that even though both the Dewdrop and An- chor lenses are in under -dense environments locally, they still require relatively large shear contributions to pro- duce good fits. In all three cases, we also note the large discrepancy between the modeled and LOS-predicted shear values. To explore this, we ran lens models with external shear and orientation restricted to the predicted values, and then examining the resulting models, and particularly the fit χ2 ν. All three new models require lenses with much higher ellipticity than the light sug- gests, though in the Dewdrop and the Anchor the formal ν remains plausible given the constraints, χ2 χ2 ν = 1.02 and 1.04 (or underfit by ∼1- and ∼2-σ), respectively. The new Cross fit, however, is strongly ruled out with χ2 ν = 2.00 (or by ∼75-σ). This suggests that at least in this case, the inferred LOS influence by SIS dark matter halos is insufficient, and that the large scale structure "sheet" must have an important additional effect. Our conclusions may be summarized as follows: 1. We have discovered two new strong galaxy-galaxy lenses by visual inspection, with reasonable lens models and source reconstructions. 2. These lenses are drawn from a range of local-density environments, which do not necessarily reflect the influence of unassociated large scale structure. 3. In at least the case of the Cross, the known large scale structure sheet at the redshift of the lens, which is not formally accounted for in the LOS calculation, has a demonstrable effect on the lens model. We thank Marusa Bradac for discussions. LAM thanks Russell Mirabelli for expert assistance with a GIMP script facilitating the inspection of the ACS data, and UC Berkeley and UC Santa Cruz for their frequent hos- pitality during the course of this work. The work of LAM was carried out at Jet Propulsion Laboratory, California Institute of Technology, under a contract with NASA. The work of PJM was supported in part by the U.S. De- partment of Energy under contract number DE-AC02- 76SF00515. JAN and ALC are supported by NASA through the Hubble Fellowship grants HF-011065.01-A and HF-01182.01-A, respectively. Auger, M. W., Fassnacht, C. D., Abrahamse, A. L., Lubin, L. M., Bar-Kana, R. 1996, ApJ, 468, 17 & Squires, G. K. 2006, astro-ph/0603448 REFERENCES Environments of gravitational lenses 5 Bolton, A. S., Burles, S., Koopmans, L. V. E., Treu, T., & Kundic, T., Hogg, D. W., Blandford, R. D., Cohen, J. G., Lubin, Moustakas, L. A. 2005, ApJL, 624, 21 -- . 2006, ApJ, 0, 0 Bolton, A. S., Burles, S., Schlegel, D. J., Eisenstein, D. J., & Brinkmann, J. 2004, AJ, 127, 1860 Coil, A. L., Newman, J. A., Kaiser, N., Davis, M., Ma, C.-P., Kocevski, D. D., & Koo, D. C. 2004a, ApJ, 617, 765 Coil, A. L., et al. 2004b, ApJ, 609, 525 Cooper, M. C., Newman, J. A., Madgwick, D. S., Gerke, B. F., L. M., & Larkin, J. E. 1997b, AJ, 114, 2276 Lupton, R., Blanton, M. R., Fekete, G., Hogg, D. W., O'Mullane, W., Szalay, A., & Wherry, N. 2004, PASP, 116, 133 Metcalf, R. B. 2005, ApJ, 629, 673 Mitchell, J. L., Keeton, C. R., Frieman, J. A., & Sheth, R. K. 2005, ApJ, 622, 81 Momcheva, I., Williams, K., Keeton, C., & Zabludoff, A. 2006, ApJ, 641, 169 Yan, R., & Davis, M. 2005a, ApJ, 634, 833 Morgan, N. D., Kochanek, C. S., Pevunova, O., & Schechter, P. L. Cooper, M. C., et al. 2005b, ArXiv Astrophysics e-prints Crampton, D., Le Fevre, O., Hammer, F., & Lilly, S. J. 1996, A&A, 307, L53 Dye, S., & Warren, S. J. 2005, ApJ, 623, 31 Fassnacht, C. D., Gal, R. R., Lubin, L. M., McKean, J. P., Squires, G. K., & Readhead, A. C. S. 2006a, ApJ, 642, 30 Fassnacht, C. D., & Lubin, L. M. 2002, AJ, 123, 627 Fassnacht, C. D., Moustakas, L. A., Casertano, S., Ferguson, H. C., Lucas, R. A., & Park, Y. 2004, ApJL, 600, L155 Fassnacht, C. D., et al. 2006b, ArXiv Astrophysics e-prints Im, M., et al. 2002, ApJ, 571, 136 Jonsson, J., Dahl´en, T., Goobar, A., Gunnarsson, C., Mortsell, E., & Lee, K. 2006, ApJ, 639, 991 Keeton, C. R., Falco, E. E., Impey, C. D., Kochanek, C. S., Leh´ar, J., McLeod, B. A., Rix, H.-W., Munoz, J. A., & Peng, C. Y. 2000, ApJ, 542, 74 Keeton, C. R., & Zabludoff, A. I. 2004, ApJ, 612, 660 Kochanek, C. S. 2004, astro-ph/0407232 Koo, D. C., et al. 1996, ApJ, 469, 535 Koopmans, L. V. E., Treu, T., Bolton, A. S., Burles, S., & 2005, AJ, 129, 2531 Ratnatunga, K. U., Ostrander, E. J., Griffiths, R. E., & Im, M. 1995, ApJL, 453, 5 Tonry, J. L. 1998, AJ, 115, 1 Tonry, J. L., & Kochanek, C. S. 1999, AJ, 117, 2034 Treu, T., Koopmans, L. V., Bolton, A. S., Burles, S., & Moustakas, L. A. 2006, ApJ, 640, 662 Treu, T., & Koopmans, L. V. E. 2004, ApJ, 611, 739 Walsh, D., Carswell, R. F., & Weymann, R. J. 1979, Nature, 279, 381 Wambsganss, J., Bode, P., & Ostriker, J. P. 2005, ApJL, 635, L1 Warren, S. J., Hewett, P. C., Lewis, G. F., Moller, P., Iovino, A., & Shaver, P. A. 1996, MNRAS, 278, 139 Wayth, R. B., Warren, S. J., Lewis, G. F., & Hewett, P. C. 2005, MNRAS, 360, 1333 Williams, K. A., Momcheva, I., Keeton, C. R., Zabludoff, A. I., & Lehar, J. 2005, astro-ph/0511593 Willis, J. P., Hewett, P. C., & Warren, S. J. 2005, MNRAS, 363, 1369 Young, P., Gunn, J. E., Oke, J. B., Westphal, J. A., & Kristian, J. Moustakas, L. A. 2006, ApJ, 0, 0 1980, ApJ, 241, 507 Kormann, R., Schneider, P., & Bartelmann, M. 1994, A&A, 284, Zepf, S. E., Moustakas, L. A., & Davis, M. 1997, ApJL, 474, L1 285 Kundic, T., Cohen, J. G., Blandford, R. D., & Lubin, L. M. 1997a, AJ, 114, 507
0805.0312
1
0805
2008-05-02T20:50:26
BD-22 5866: A Low-mass Quadruple-lined Spectroscopic AND Eclipsing Binary
[ "astro-ph" ]
We report our discovery of an extremely rare, low mass, quadruple-lined spectroscopic binary BD-22 5866 (=NLTT 53279, integrated spectral type = M0 V), found during an ongoing search for the youngest M dwarfs in the solar neighborhood. From the cross-correlation function, we are able to measure relative flux levels, estimate the spectral types of the components, and set upper limits on the orbital periods and separations. The resulting system is hierarchical composed of K7 + K7 binary and a M1 + M2 binary with semi-major axes of asini_{A}<=0.06 AU and asini_{B}<=0.30 AU. A subsequent search of the SuperWASP photometric database revealed that the K7 + K7 binary is eclipsing with a period of 2.21 days and at an inclination angle of 85 degrees. Within uncertainties of 5%, the masses and radii of both components appear to be equal (0.59 Msun, 0.61 Rsun). These two tightly orbiting stars (a = 0.035 AU) are in synchronous rotation causing the observed excess Ca II, Halpha, X-ray and UV emission. The fact that the system was unresolved with published adaptive optics imaging, limits the projected physical separation of the two binaries at the time of the observation to d_{AB} < 4.1 AU at the photometric distance of 51 pc. The maximum observed radial velocity difference between the A and B binaries limits the orbit to asini_{AB}<=6.1 AU. As this tight configuration is difficult to reproduce with current formation models of multiple systems, we speculate that an early dynamical process reduced the size of the system such as the interaction of the two binaries with a circumquadruple disk. Intensive photometric, spectroscopic and interferometric monitoring as well as a parallax measurement of this rare quadruple system is certainly warranted.
astro-ph
astro-ph
BD -- 22◦5866: A Low-mass Quadruple-lined Spectroscopic and Eclipsing Binary1 Evgenya Shkolnik2 NASA Astrobiology Institute, Institute for Astronomy, University of Hawaii at Manoa 2680 Woodlawn Drive, Honolulu, HI 96822 [email protected] Michael C. Liu3 Institute for Astronomy, University of Hawaii at Manoa 2680 Woodlawn Drive, Honolulu, HI 96822 [email protected] I. Neill Reid Space Telescope Science Institute, Baltimore, MD 21218 [email protected] Leslie Hebb, Andrew C. Cameron School of Physics and Astronomy, University of St. Andrews, North Haugh St Andrews, Fife Scotland KY16 9SS [email protected], [email protected] Laborat´orio Nacional de Astrof´ısica/MCT, Rua Estados Unidos 154, 37504-364 Itajub´a, Carlos A. Torres Brazil [email protected] and David M. Wilson Astrophysics Group, Keele University, Staffordshire, ST5 5BG [email protected] -- 2 -- ABSTRACT We report our discovery of an extremely rare, low mass, quadruple-lined spec- troscopic binary BD -- 22◦5866 (=NLTT 53279, integrated spectral type = M0 V), found during an ongoing search for the youngest M dwarfs in the solar neighbor- hood. From the cross-correlation function, we are able to measure relative flux levels, estimate the spectral types of the components, and set upper limits on the orbital periods and separations. The resulting system is hierarchical composed of K7 + K7 binary and a M1 + M2 binary with semi-major axes of aAsiniA≤0.06 AU and aBsiniB≤0.30 AU. A subsequent search of the SuperWASP photometric database revealed that the K7 + K7 binary is eclipsing with a period of 2.21 days and at an inclination angle of 85◦. Within uncertainties of 5%, the masses and radii of both components appear to be equal (0.59 M⊙, 0.61 R⊙). These two tightly orbiting stars (a = 0.035 AU) are in synchronous rotation causing the observed excess Ca II, Hα, X-ray and UV emission. The fact that the sys- tem was unresolved with published adaptive optics imaging, limits the projected physical separation of the two binaries at the time of the observation to dAB.4.1 AU at the photometric distance of 51 pc. The maximum observed radial velocity difference between the A and B binaries limits the orbit to aABsiniAB≤6.1 AU. As this tight configuration is difficult to reproduce with current formation models of multiple systems, we speculate that an early dynamical process reduced the size of the system such as the interaction of the two binaries with a circumquadruple disk. Intensive photometric, spectroscopic and interferometric monitoring as well as a parallax measurement of this rare quadruple system is certainly warranted. Subject headings: binaries: spectroscopic, eclipsing, stars: low-mass, stars: individual: BD -- 22◦5866 late-type, activity, 1Based on observations collected at the W. M. Keck Observatory and the Canada-France-Hawaii Tele- scope. The Keck Observatory is operated as a scientific partnership between the California Institute of Technology, the University of California, and NASA, and was made possible by the generous financial sup- port of the W. M. Keck Foundation. The CFHT is operated by the National Research Council of Canada, the Centre National de la Recherche Scientifique of France, and the University of Hawaii. 2NASA Postdoctoral Fellow 3Alfred P. Sloan Research Fellow -- 3 -- 1. Introduction The multiplicity of stars is an important constraint of star formation theories as most stars form as part of a binary or higher-order multiple system (Duquennoy & Mayor 1991, Fischer & Marcy 1992, Halbwachs et al. 2003). Moreover, double- (or multi-) lined spectro- scopic binaries (SBs) allow precise determination of dynamical properties including the mass ratio, and if the inclination can be determined, the individual component masses. And given their common age and metallicity, SBs are very useful in calibrating stellar evolutionary models. Here, we report our detection of a visually unresolved quadruple-lined spectroscopic and eclipsing (ESB4) binary BD -- 22◦5866 (=NLTT 53279; V = 10.1, J2MASS = 7.54) composed of four low-mass stars. As part of our search for young M dwarfs within 25 pc (Shkolnik et al. 2006), we acquire high-resolution spectra of cool dwarfs compiled by the NStars project (Reid et al. 2003, 2004) that have X-ray luminosities comparable to or greater than Pleiades members. We cross-correlate these spectra with radial velocity (RV) standard stars in order to determine the RVs needed to measure galactic space motion. This process is sensitive to finding SBs, particularly those in short-period orbits, while the sample in general is biased towards tidally-locked systems whose rapid rotation produces high chromospheric and X-ray emission. BD -- 22◦5866 was on our list of candidate young M dwarfs because of its high fractional X-ray1 and UV2 flux (fX/fJ = 1.94×10−3, fN U V /fJ = 1.6×10−4, fF U V /fJ = 3×10−5) and photometrically determined distance of 20.7±3.4 pc (Reid et al. 2004). In this paper we decompose BD -- 22◦5866 into a tight hierarchical quadruple system consisting of an eclipsing K7 + K7 binary and a M1 + M2 binary, and discuss the implications of this extremely unusual multiple system for the dynamical evolution of binary stars. BD -- 22◦5866 must be quite rare since we detected only one such SB4 in our sample of 196 X-ray-selected M dwarfs. Other high-resolution spectroscopic surveys for SB's have turned up very few if any tight SB4s; e.g. Udry et al. (1998) surveyed 3347 G dwarfs and found none, while Torres et al. (2006) flagged two SB4's in a X-ray-selected sample of 1511 stars with B − V > 0.6. Of these two, one is BD -- 22◦5866 (independently discovered) and the other (CPD-64 4353) is a higher mass system whose orbital limits are unconstrained. These limit the frequencies of SB4s to 2 in 4000 -- 5000, with only BD -- 22◦5866 consisting of four low-mass stars. Note that Torres et al.'s and our samples are biased towards active close-in SBs with strong X-ray emission and high orbital velocities more easily resolved with the 1X-ray data is from the ROSAT All-Sky Survey Bright Source Catalogue (Vogues et al. 1999). 2Near and far UV (NUV, FUV) fluxes were measured by the GALEX (GALaxy Evolution eXplorer) All-Sky Survey (Morrissey et al. 2007). -- 4 -- CCF, making the intrinsic frequency of these types of systems even lower. Though there are a couple of dozen hierarchical quadruples listed in the literature, they mostly consist of two visually resolved SB2s at wide (dAB &30 AU) separations, with a handful of them exhibiting four sets of spectral lines. (See Torres et al. 2007 and references therein.) We are aware of only one previously known spatially unresolved quadruple-lined spectroscopic binary: XY Leo, a W UMa-type binary (Barden 1987, Pribulla et al. 2007), making BD -- 22◦5866 only the second such system studied and the first to be comprised of only low-mass stars.3 The system also contains the thirteenth published eclipsing binary (EB) with component masses less than 0.7 M⊙, and the only one known to be in a quadruple system. 2. The spectra We acquired 4 high-resolution ´echelle spectra of BD -- 22◦5866, 2 with the High Reso- lution Echelle Spectrometer (HIRES; Vogt et al. 1994) on the Keck I 10-m telescope and 2 with the ´Echelle SpectroPolarimetric Device for the Observation of Stars (ESPaDOnS; Do- nati et al. 2006) on the Canada-France-Hawaii 3.6-m telescope, both located on the summit of Mauna Kea. We used the 0.861′′ slit with HIRES to give a spectral resolution of λ/∆λ≈58,000. The upgraded detector consists of a mosaic of three 2048 x 4096 15-µm pixel CCDs, corresponding to a blue, green and red chip spanning 4900 -- 9300 A. To maximize the throughput near the peak of a M dwarf spectral energy distribution, we used the GG475 filter with the red cross-disperser. The data product of each exposure is a multiple-extension FITS file from which we reduce and extract the data from each chip separately. ESPaDOnS is fiber fed from the cassegrain to coud´e focus where the fiber image is projected onto a Bowen-Walraven slicer at the spectrograph entrance. With a 2048×4608- pixel CCD detector, ESPaDOnS' 'star+sky' mode records the full spectrum over 40 grating orders covering 3700 to 10400 A at a spectral resolution of λ/∆λ≈68,000. The data were reduced using Libre Esprit, a fully automated reduction package provided for the instrument and described in detail by Donati et al. (1997, 2007). Each stellar exposure is bias-subtracted and flat-fielded for pixel-to-pixel sensitivity 3GG Tau is a famous pre-main sequence low-mass quadruple system but in a much wider orbital config- uration than BD −22◦5866. The distance between binaries A and B is 1414 AU with the Aa + Ab and Ba + Bb separated by 35 and 207 AU, respectively (White et al. 1999). -- 5 -- variations. After optimal extraction, the 1-D spectra are wavelength calibrated with a Th/Ar arc. Finally the spectra are divided by a flat-field response and corrected for the heliocentric velocity. The final spectra were of moderate S/N reaching ≈ 70 per pixel at 7000 A. Each night, spectra were also taken of an A0V standard star for telluric line correction and an early-M radial velocity standard. A log of the observations is presented in Table 1, and a portion of a spectrum is shown in Figure 1. The integrated spectral type (SpT) was measured to be M0 from the ratio of the band indices of TiO λ7140 to TiO λ8465 defined by Slesnick et al. (2006). This is slighter earlier than the M0.5 determined from low resolution spectra using the TiO λ7140 index by Reid et al. (2004). We cross-correlated each of 7 orders between 7000 and 9000 A of each stellar spectrum with a RV standard of similar spectral type using IRAF's1 fxcor routine (Fitzpatrick 1993). We excluded the Ca II infrared triplet (IRT)4 and regions of strong telluric absorption in the cross-correlation and removed a broad, low order curvature from each CCF which varies from order to order and is due to differences in the continuum between BD -- 22◦5866 and the RV templates. A representative cross-correlation function (CCF) for each of the four observations is shown in Figure 2. We were able to clearly resolve the CCF peaks of all four stellar components on May 11, 2006. This allowed us to estimate the spectral types of the four stars assuming a flux- weighted relation (Cruz & Reid 2002) between component and integrated spectral types: SpTint = (fAaSpTAa + fAbSpTAb + fBaSpTBa + fBbSpTBb)/(fAa + fAb + fBa + fBb), where fAa, fAb, fBa and fBb are derived from the integrated flux of the gaussians fits to the four cross- correlation peaks. In this case, fAa ≈ fAb and fBa ≈ fBb ≈ 0.3fAa such that the difference in I magnitude between Aa and Ba is 1.2 mag. To measure the error associated with these relative fluxes, we cross-correlated the spectrum with three template spectra of stars with low vsini and differing SpTs. These were GJ 2079 (K7, v=4.1 km s−1 as calculated from Prot of 7.78 d; Pizzolato et al. 2003), GJ 908 (M1, vsini=3 km s−1; Glebocki et al. 2000), and GJ 436 (M2.5, vsini≤1 km s−1; Glebocki et al. 2000). The average results give fAa : fAb : fBa : fBb = 1 : 0.99 : 0.33 : 0.23. These agree to better than 15%, with the bulk of the error lying with fBb whose integrated fluxes ranged from 0.2 to 0.3 of fAa. 1IRAF (Image Reduction and Analysis Facility) is distributed by the National Optical Astronomy Ob- servatories, which is operated by the Association of Universities for Research in Astronomy, Inc. (AURA) under cooperative agreement with the National Science Foundation. 4Since the Ca II emission was clearly stronger in two of the four stars and potentially variable, it was important to eliminate the IRT from the cross-correlation. -- 6 -- From the spectra alone, we estimated the component spectral types to be K7 + K7 and M1 + M2. These spectral types are consistent with the distinctly redder location of BD -- 22◦5866 on a color-color diagram than expected if it were a single M0 dwarf. For example, it lies ∼0.7 mags redder than expected for an M0 dwarf on a (V − J)-(I − J) plot (Reid & Hawley 2005). BD -- 22◦5866 is 3.7 times more luminous than a single M0 dwarf, and correcting for this over-luminosity, the system's photometric distance is at about 51 pc. We measure the RVs (listed in Table 1) of each component from the gaussian peaks fitted to the CCFs, taking the average of all orders, with a RMS typically less than 1 km s−1. Severe blending of the CCF peaks resulted in poor gaussians fits to the two latest observations. One additional set of RVs was extracted from a low-S/N spectrum collected using FEROS on La Silla's 1.5-m telescope. The instrument and data reduction of this measurement is presented in Torres et al. 2006. 3. The light curve With the high eclipsing probability (> 10%) of the tight Aa+Ab binary, we searched for existing photometry of this system in the database of the UK Wide-Angle Search for Planets (WASP) project. The WASP project is a wide-angle, robotic photometric monitor- ing campaign designed to detect significant numbers of transiting extrasolar planets using a single pass band from 4000 to 7000 A. The project instrumentation, infrastructure, and data processing are are described in detail in Pollacco et al. (2006). In short, SuperWASP (La Palma, Canary Islands, Spain) and WASP-South (Sutherland, South Africa) instruments each consist of 8 small aperture camera lenses (8 × 200mm, f/1.8) backed by wide-format CCDs with a plate scale of 14.2′′/pixel. The cameras are attached to a single robotic mount such that the field-of-view of a single WASP pointing is approximately 15◦ × 30◦. The mount scans the sky taking repeated individual exposures in a single declination band over the ob- servable sky every clear night to obtain a time-series of images with typical sampling rate of ∼ 8 minutes. The raw images are processed via a custom-built pipeline with aperture photometry performed using a 3.5 pixel aperture (48′′) on all observed stars. The resulting differential photometry light curves are searched for periodic dips in brightness using a mod- ified box-least squares search algorithm which is outlined in Collier Cameron et al. (2006). The high-cadence, high precision time-series photometry produced for millions of bright stars as part the WASP transit survey is extremely good for detecting and characterizing eclipsing binaries in addition to the primary science goal of transiting planet detection. The WASP-South data set on BD -- 22◦5866 consists of 9531 photometric measurements obtained over two observing seasons (May 2006 -- October 2007). The observed light curve shows -- 7 -- periodic (P =2.21 days), nearly equal depth (≈ 0.2 mag) primary and secondary eclipses identifying the tight K7 + K7 pair as an eclipsing binary. The raw WASP light curve is shown in Figure 3 phased according to the eclipse ephemeris (see Table 2). In addition to the eclipses, the photometry also exhibits modulating out-of-eclipse variability with an amplitude of 0.01 -- 0.03 mags which is likely caused by asymmetric starspots on the surface of the individual binary components. The measured period of this variability as determined from the strongest peak of the periodogram (Figure 4) is nearly identical to the binary orbital period and tracks the rotation of the star indicating the system is tidally synchronized. This is expected since the synchronization time-scale for the short-period binary A is only 70 Myr (Zahn 1977), several orders of magnitude shorter than for binary B, making it the likely source of the observed X-ray and UV flux. The RVs also confirm that Aa and Ab are the sources of the moderate emission we see in the lines of the Ca II IRT (Figure 1) and Hα (EWHα(Aa)= -- 0.37 ± 0.01 A, EWHα(Ab)= -- 0.55 ± 0.01 A). We therefore conclude the system is not particularly young, consistent with the fact that we see no sign of Li (λ6708A, EWLi < 21 mA). Comparing these characteristics of binary A with the late-type systems of the Strassmeier et al. 2003 catalog of chromospherically active binary stars, we can classify BD -- 22◦5866 A as BY Dra binary. Due to the eclipsing nature of the K7 + K7 binary, the individual stellar radii, tem- perature ratio and orbital inclination can be measured directly from the light curve. We first remove the out-of-eclipse variability before modeling the eclipsing system and deriving the physical properties of the individual stars. Because the amplitude of the modulation is changing slowly over the course of the observations, we separate the photometric data into 3-week time bins and remove the variability of each segment independently. For each time segment, we solve for the coefficients of a sine curve that best fits the out-of-eclipse data using least-squares minimization and then subtract this model from the observed photom- etry. We then derive the properties of the system using an implementation of the EBOP eclipsing binary modeling code (Southworth et al. 2004, Popper & Etzel 1981). The recti- fied light curve around the primary and secondary eclipses is shown in Figure 5 with the best fitting model over-plotted. The results of the analysis give individual stellar radii of RAa = 0.614 ± 0.045R⊙ and RAb = 0.598 ± 0.045R⊙ (using the semi-major axis5 derived from the RV curve, see Figure 6 and Table 2), an orbital inclination of 85.5◦ ± 1◦, a temperature ratio TAb/TAa = 0.98 ± 0.02, and an eccentricity of 0.00 ±0.01. 5We have used the luminosity ratio of stellar components Aa and Ab derived from the cross-correlation functions (See Section 2) as an independent constraint on the ratio of the radii and the amplitude of the radial velocity curve described in the following section to derive the orbital separation. -- 8 -- 4. The orbital configuration Although there are not many RV measurements, we can determine well-constrained masses due to the accurate ephemeris derived from the photometry. With the precise eclipse period, we are able to plot our RV measurements to determine the remaining orbital parame- ters for the K7 + K7 binary. To reduce the number of fitted parameters needed to derive the component masses, we adopt a mass ratio of 1.0 for the binary because the two components have nearly the same flux and temperature values. We solved for a single RV amplitude and systemic velocity γ which best fit the Keck and CFHT RVs outside of eclipse. The 5-year span between the Mauna Kea and the La Silla observations is a significant fraction of the A + B orbital period, and the corresponding shift in γ due to the presence of binary B is apparent in the offset of those points in Figure 6. The true systemic velocity of the entire system is close to -- 9 km s−1 measured by taking the average all observed RVs. The mass for each K7 dwarf derived from the RV curve is 0.59 M⊙. We estimate the uncertainty on this mass to be ≈5% based on the potential deviations from a mass ratio of 1. Additional RV measurements will allow us to derive component masses with <1% accuracy. The relative flux levels of the CCF peaks indicate that these two stars are in between a K7 and M0 dwarf, consistent with their derived masses. The system parameters are summarized in Table 2. This pair is only the thirteenth confirmed EB with component masses less than 0.7 M⊙. Having model-independent masses and radii for both components is an essential part of testing stellar evolution models (Baraffe et al. 1998). Although the measured properties of the K7 binary components are consistent with being on the main sequence, the two stars appear ≈10% larger in radii for their masses than predicted by the stellar evolution models (Figure 7), in agreement with recent measurements of the other low-mass EBs with 10-20% larger radii than expected. (See Table 3) Some empirical data suggest this is due to the effects of magnetic activity (L´opez-Morales 2007) while other data show no correlation with activity, and instead find metallicity to be the dominant cause (Berger et al. 2006). With regards to the rest of the quadruple system, the maximum velocity separation measured for the lower-mass pair (Ba + Bb) is 52 ± 0.8 km s−1. This coupled with the estimated spectral types (and masses 0.49 and 0.44 M⊙ for a M1 and M2 dwarf, respectively; Reid & Hawley 2005), allows us to set upper limits on its orbital period and semi-major axis of 62 days and 0.30 AU. Though there is no indication of eclipses for this wider binary, we cannot determine if Ba and Bb are in a coplanar orbit with binary A. With a grazing eclipse angle of >89◦, Ba + Bb would not exhibit eclipses even if coplanar with Aa + Ab. On 11 May, 2006, we observed a maximum relative orbital velocity of 17.5 ± 2.3 km s−1 -- 9 -- for the A and B binaries. If we treat the system as two single objects in Keplerian orbit, then this velocity measurement sets an upper limit of aABsiniAB≤ 6.1 AU and PAB ≤ 3764 d. At this separation, the binary pair A + B is resolvable with ground-based adaptive optics imaging.6 Though the change in γ of ≈17 km s−1 in the five years between observations implies that the period is likely closer to 6 years (and then aABsiniAB ∼ 4 AU) rather than the 10 years of the maximal orbit. A schematic of this ESB4 with its limiting orbital parameters is shown in Figure 8. In order to use the empirical criterion for the orbital stability of binary systems (Eggleton & Kiseleva 1995), we approximate this system to be a hierarchical triple by treating the tight Aa+Ab binary as a single mass, and conclude that even for high values of eccentricity, BD -- 22◦5866 is very stable. 5. A need for a primordial circumquadruple disk? In order to reproduce the typical observed separations of ∼40 AU in binary systems, Sterzik et al. (2003) determined that a multiple star system must undergo a phase of dynam- ical evolution after the fragmentation stemming from isothermal collapse of the molecular cloud. However, the formation of very close binaries with a on the order of 1 AU, as is the case for BD -- 22◦5866 AB, is still somewhat unclear. Tokovinin et al.'s (2006) determination that the vast majority of short-period binaries must have tertiary components with which angular momentum can be exchanged may explain the tight orbit of Aa+Ab, yet also implies that aAB must have been even smaller in the past. This same argument may imply that a distant fifth component to the system might have interacted with binaries A and B to bring them to their small separation. However, no common proper motion companions were found within 2000 AU of the system.7 We speculate that an earlier dynamical process reduced the physical size of the system, such as binary-disk interactions within a circumquadruple disk. This is supported by the simulations by Bate et al. (2002) of binary- and triple-star formation which predict the shrinking of orbital distances to .10 AU through accretion and 6Daemgen et al. (2007) observed BD -- 22◦5866 to have no physical companions within 0.08′′ using adaptive optics, setting an upper limit on the physical separation between A and B of dAB < 4.1 AU at 51 pc from Earth at the observed orbital phase. 7We searched the Naval Observatory Merged Astrometric Dataset (NOMAD; Zacharias et al. 2005) which compiles data from several catalogs including the 2MASS catalog (Skrutskie et al. 2006). With a K-band magnitude limit of 16, MK is limited to 12.4 mags at the system distance of 51 pc. This sets the maximum mass of any potential companion to 20 -- 50 MJ , the mass of an old field mid- to late-L dwarf. -- 10 -- interactions with circumbinary and circumtriple gas-rich disks on the time-scale of less than 105 years. Though in the case of a hierarchical triple, the tertiary component is always in a wide orbit of more than 30 AU from the tight binary. They also calculate that the specific angular momentum exchange will drive the masses of the two components towards equality which is marginally the case for the A and B binaries of BD -- 22◦5866 which has a mass ratio q ≈ 0.7.8 There is no theoretical discussion in the literature regarding circumquadruple disks around low-mass stars. Yet it is reasonable to assume that BD -- 22◦5866 must have once had such a gas-rich primordial disk, which likely lasted for less than 105 -- 106 years (Artymowicz & Lubow 1994). This would imply that the components accumulated their masses and interacted dynamically with each other (Sterzik et al. 2005) and their disk during the same early stages of their evolution. A systematic interferometric investigation of young pre- main sequence stars to search for additional low-mass quadruple systems will help test our hypothesis of stellar interactions with a circumquadruple disk, since these very young systems should be wider. 6. Future observations Since the BD -- 22◦5866 system is both bright and contains short orbital periods, we encourage those actively monitoring binaries to include this low-mass system in their pro- grams. Spectroscopic monitoring will yield the component velocities for which specially designed techniques, such as the broadening function formulism of Rucinski et al. (2002) or the four-dimensional cross-correlation method by Torres et al. (2007), are well suited. These are necessary to determine the Keplerian orbital parameters, which will provide more accurate mass ratios, and if the two remaining inclinations can be measured, the masses of the individual components, arguably the most important stellar property in the context stellar evolution (e.g., interferometric observations (i.e. Boden et al. 2005) offer very high spatial resolution and could measure the separation and inclination of the A+B system.) In addition, a parallax measurement to determine a more accurate distance to the system would be very beneficial. With the potential of these observations to yield all four stellar masses as well as the three orbital inclinations, the BD -- 22◦5866 system could provide key constraints to possible 8This may also explain the even tighter, higher mass system of VW LMi consisting of one non-eclipsing detached binary and one eclipsing contact binary at a separation of dAB = 1.24 AU and with solar-mass components (Pribulla et al. 2006). -- 11 -- formation scenarios, as well as the ability to test the hypothesis that orbital evolution within a disk would create coplanar orbits (Bonnell & Bate 1994). Clearly, intensive follow-up observations of this rare eclipsing SB4 are warranted. E.S appreciates useful discussion with George Herbig, Slavek Rucinski, Katelyn Allers and Nader Haghighipour, as well as helpful suggestions from the anonymous referee. Also, thank you to Nick Dunstone for his quick and independent discovery of the eclipses in the WASP database and the CFHT and Keck staff for their care in setting up the instruments and support in the control room. Research funding from the NASA Postdoctoral Program (formerly the NRC Research Associateship) for E.S. is gratefully acknowledged. This mate- rial is based upon work supported by the National Aeronautics and Space Administration through the NASA Astrobiology Institute and the NASA/GALEX grant program under Co- operative Agreement Nos. NNA04CC08A and NNX07AJ43G issued through the Office of Space Science. M.C.L. acknowledges support from the Alfred P. Sloan Research Fellowship. Artymowicz, P., Lubow, S. H., 1994, ApJ, 421, 651 REFERENCES Baraffe, I.; Chabrier, G.; Allard, F.; Hauschildt, P. H., 1998, A&A, 33, 403 Barden, S.C., 1987, ApJ, 317, 333 Bate, M.R.; Bonnell, I.A. I.; Akeson, R. L.; Carpenter, J. M.; Torres, G.; Latham, D. W.; Soderblom, D.R.; Nelan, E.; Franz, O. G.; Wasserman, L.H., 2005, ApJ, 635, 442 Bayless, A. J., & Orosz, J. A. 2006, ApJ, 651, 1155 Becker, A.C. et al., MNRAS, in press (astroph:0801.4474v1) Berger, D. H.; Gies, D. R.; McAlister, H. A.; ten Brummelaar, T. A.; Henry, T. J.; Sturmann, J.; Sturmann, L.; Turner, N. H.; Ridgway, S. T.; Aufdenberg, J. P.; M´erand, A., 2006, ApJ, 644, 475 Blake, C. H., Torres, G., Bloom, J. S., & Gaudi, B. S. 2007, ArXiv e-prints, 707, arXiv:0707.3604 Boden, A. F., et al. 2005, ApJ, 635, 442 Bonnell, I., Bate, M. R. 1994, MNRAS, 269, L45 -- 12 -- Butler, R. P., Wright, J. T., Marcy, G. W., Fischer, D. A., Vogt, S. S., Tinney, C. G., Jones, H. R. A., Carter, B. D., Johnson, J. A., McCarthy, C., Penny, A. J., 2006, ApJ, 646, 505 Collier Cameron, A., et al. 2006, MNRAS, 373, 799 Creevey, O. L., et al. 2005, ApJ, 625, L127 Cruz, K. L., & Reid, I. N. 2002, AJ, 123, 2828 Daemgen, S., Siegler, N., Reid, I.N., Close, L.M., 2007, ApJ, 654, 558 Delfosse, X., Forveille, T., Mayor, M., Burnet, M., & Perrier, C. 1999, A&A, 341, L63 Donati, J.-F.; Catala, C.; Landstreet, J. D.; Petit, P., 2006, ASPC, 358, 362 Duquennoy, A., Mayor, M. 1991, A&A, 248, 485 Eggleton, P., Kiseleva, L., 1995, ApJ, 455, 640 Fischer, D. A., & Marcy, G. W., 1992, ApJ, 396, 178 Fitzpatrick, M. J. 1993, in ASP Conf. Ser. 52, Astronomical Data Analysis Software and Systems II, ed. R. J. Hanisch, R. V. J. Brissenden, & J. Barnes (San Francisco: ASP), 472 Glebocki R., Gnacinski P., Stawikowski A., 2000, AcA, 50, 509 Halbwachs, J.-L., Mayor, M., Udry, S., Arenou, F. 2003, A&A, 397, 159 Hebb, L., Wyse, R. F. G., Gilmore, G., & Holtzman, J. 2006, AJ, 131, 555 L´opez-Morales, M., 2007, ApJ, 660, 732 L´opez-Morales, M., & Shaw, J. S. 2007, The Seventh Pacific Rim Conference on Stellar Astrophysics, 362, 26 Maceroni, C., & Montalb´an, J. 2004, A&A, 426, 577 Metcalfe, T. S., Mathieu, R. D., Latham, D. W., & Torres, G. 1996, ApJ, 456, 356 Morrissey, P. et al., 2007, ApJS, 173, 682 Mullan, D.J., MacDonald, J., 2001, ApJ, 559, 353 Pizzolato N., Maggio A., Micela G., Sciortino S., Ventura P., 2003, A&A, 397, 147 -- 13 -- Pollacco, D. L., et al. 2006, PASP, 118, 1407 Popper, D.M & Etzel, P.B., 1981, AJ, 86, 102 Pribulla, T., Rucinski, S. M., Lu, W., Mochnacki, S. W., Conidis, G., Blake, R. M., DeBond, H., Thomson, J. R., Pych, W., Ogloza, W., Siwak, M., 2006, AJ, 132, 769 Pribulla, T., Rucinski, S. M., Conidis, G., DeBond, H., Thomson, J. R., Gazeas, K., Ogloza, W., 2007, AJ, 133, 1977 Reid, I. N., Gizis, J. E., Hawley, S. L. 2002, AJ, 124, 2721 Reid, I. N., et al. 2004, AJ, 128, 463 Reid, I. N., Hawley, S. L., New Light on Dark Stars Red Dwarfs, Low-Mass Stars, Brown Stars, 2005, Praxis Publishing Ltd, ISBN 3-540-25124-3 Rucinski, S. M., 2002, AJ, 124, 1746 S´egransan, D., Delfosse, X., Forveille, T., Beuzit, J.-L., Udry, S., Perrier, C., & Mayor, M. 2000, A&A, 364, 665 Skrutskie, M.F. et al., 2006, AJ, 131, 1163 Slesnick, C. L., Carpenter, J. M., Hillenbrand, L. A., 2006, AJ, 131, 3016 Southworth, J., Maxted, P.F.L., Smalley, B., 2004, MNRAS, 349, 547 Sterzik, M. F., Durisen, R. H., Zinnecker, H., 2003, A&A, 411, 91 Sterzik, M. F., Melo, C. H. F., Tokovinin, A. A., van der Bliek, N, 2005, A&A, 434, 671 Strassmeier K.G., Hall D.S., Fekel F.C., and Scheck M., 1993, A&AS, 100, 173 Tokovinin, A., Thomas, S., Sterzik, M., Udry, S., 2006, A&A, 450, 681 Torres, G, Neuhuser, R, Guenther, E. W., 2002, AJ, 123, 1701 Torres C.A.O., Quast G.R., da Silva L., de la Reza R., Melo C.H.F., 2006, A&A, 460, 695 Torres, G., Latham, D. W., Stefanik, R. P., 2007, ApJ, 662, 602 Udry, S., Mayor, M., Latham, D. W., Stefanik, R. P., Torres, G., Mazeh, T., Goldberg, D., Andersen, J., Nordstrom, B., 1998, ASPC, 154, 2148 Voges W. et al. , 1999, A&A, 349, 389 -- 14 -- White, Russel J., Ghez, A. M., Reid, I. N., Schultz, G., 1999, ApJ, 520, 811 Young, T. B., Hidas, M. G., Webb, J. K., Ashley, M. C. B., Christiansen, J. L., Derekas, A., & Nutto, C. 2006, MNRAS, 370, 1529 Zacharias N., Monet D.G., Levine S.E., Urban S.E., Gaume R., Wycoff G.L., 2005, AAS, 205, 4815 Zahn, J.-P., 1977, A&A, 57, 383 This preprint was prepared with the AAS LATEX macros v5.2. -- 15 -- Table 1. Observations & Radial Velocities Date Instrument HJD-2450000 φAa+Ab RVa of Aa, Ab, Ba, Bb (km s−1) RV STD, SpTb 11-May-06 12-Aug-06 4-Oct-06 5-Jul-07 Keck I/HIRES Keck I/HIRES CFHT/ESPaDOnS CFHT/ESPaDOnS 3867.12294 3959.97821 4013.83591 4287.99493 0.12994 −69.0, +58.4, −40.2, −5.4 0.12549 −66.9, +57.3, + 7 .5 , +7 .5 0.48365 −10 .2 , +0 .5 , −53.2, +0.3 0.47725 −18 .0 , −18 .0 , −25.9, +8.5 GJ 436c , M2.5 GJ 908, M1 GJ 205, M1.5 GJ 821, M1 31-Aug-01d FEROS/1.5m 2152.80326 0.79647 +64.5, −109.1, 3.7, −15.0 τ Ceti, G8.5 aFor the Keck and CFHT data, heliocentric RVs are determined by taking the average of the RVs measured individually from 7 spectral orders with RMS .1 km s−1. Values in italics represent those from CCF peaks that are strongly blended. bThe RVs for these standards are published by Marcy & Benitz (1989), except GJ 821, which is from Nidever et al. 2002. cThe velocity amplitude of the Neptune-mass planet orbiting GJ 436 is only 0.019 km s−1 (Butler et al. 2006), negligible for our purposes. dDetails of this spectrum can be found in Torres et al. 2006. Table 2. Parameters for Binary Aa + Ab Quantity Value Error Period (days) Epoch (HJD − 2450000) i (degrees) e ω RAa/a RAb/a Tef f,Aa/Tef f,Ab MAa, MAb (M⊙) a (AU) RAa (R⊙) RAa (R⊙) 2.21107(5) 3937.5900(0) 0.000004 0.00041 85.5 0.00 82a 0.0814 0.0792 0.98 0.5881 0.0351 0.614 0.598 1.0 0.01 -- 0.0060 0.0058 0.02 0.029 0.0024 0.045 0.045 aWith an eccentricity of 0, the angle of periastron ω is very poorly constrained. -- 16 -- Table 3. Known low-mass EBs Name Mass M⊙ δM M⊙ R R⊙ δR R⊙ Ref. 0.730 0.693 RXJ0239.1 A RXJ0239.1 B 2MASS J01542930+0053266 A 0.66 2MASS J01542930+0053266 B 0.62 GU Boo A GU Boo B YY Gem A YY Gem B BD -- 22◦5866 Aa BD -- 22◦5866 Ab NSVS01031772 A NSVS01031772 B UNSW-TR-2 A UNSW-TR-2 B Tres-Her0-07621 A Tres-Her0-07621 B 2MASS 04463285+1901432 A 2MASS 04463285+1901432 B OGLE BW3 V38 A OGLE BW3 V38 B CU Cnc A CU Cnc B SDSS-MEB-1 A SDSS-MEB-1 B CM Dra A CM Dra B 0.610 0.599 0.599 0.599 0.588 0.588 0.5428 0.4982 0.529 0.512 0.493 0.489 0.47 0.19 0.44 0.41 0.433 0.389 0.272 0.240 0.2307 0.2136 0.009 0.006 0.03 0.03 0.007 0.006 0.0047 0.0047 0.029 0.029 0.0027 0.0025 0.035 0.035 0.003 0.003 0.05 0.02 0.07 0.09 0.0017 0.0014 0.02 0.022 0.001 0.001 0.741 0.703 0.64 0.61 0.623 0.62 0.6191 0.6191 0.614 0.598 0.526 0.5088 0.641 0.608 0.453 0.452 0.56 0.21 0.51 0.44 0.4317 0.3908 0.268 0.248 0.2516 0.2347 L´opez-Morales & Shaw 2007 L´opez-Morales & Ribas 2005 Creevey 2005 Hebb et al. 2006 Maceroni & Montalb´an 2004 this work Young et al. 2006 Becker et al. 2008 0.004 0.002 0.08 0.09 0.016 0.02 0.0057 Delfosse et al. 1999 0.0057 0.045 0.045 0.0028 L´opez-Morales 2006 0.003 0.05 0.06 0.06 0.06 0.02 0.01 0.04 0.06 0.0052 Delfosse et al. 1999 0.0094 0.009 0.0084 0.002 Metcalf et al. 1996 0.0019 Blake et al. 2007 -- 17 -- y t i s n e t n I d e z i l a m r o N 1 0.9 0.8 0.7 0.6 1 0.8 0.6 0.4 0.2 Ab Aa BD -22°5866 Ba Bb Aa Ab GJ 436 Ca II Fe I 8650 8660 8670 8680 8690 8700 Wavelength (Å) Fig. 1. -- An example spectrum of BD -- 22◦5866 and RV standard GJ 436 observed on May 11, 2006. The spectral lines of the four components are marked by the vertical dashes. The strong Ca II emission from the Aa and Ab components is apparent at 8660 and 8664 A. -- 18 -- 11 May 2006 12 Aug 2006 04 Oct 2006 05 Jul 2007 0.4 0.3 0.2 0.1 0 0.5 0.4 0.3 0.2 0.1 0.4 0.3 0.2 0.1 0 0.25 0.2 0.15 0.1 0.05 0 -150 -100 -50 50 Radial Velocity (km/s) 0 100 150 Fig. 2. -- The cross-correlation function (with baseline removed) for BD -- 22◦5866 as mea- sured against a RV standard star on each of the four nights. The dashed curves are the gaussians fitted to the CCF peaks. -- 19 -- Fig. 3. -- The raw WASP-S light curve of BD -- 22◦5866 phased on the eclipse period of 2.21 days. The difference in starspot intensity from the two epochs is clear in the double-lined region of the light curve. -- 20 -- Fig. 4. -- Periodogram of the sine-like out-of-eclipse variability in the WASP-S light curve. We tested 5000 frequencies between 0.4 -- 20 days and found the parameters of a sine-curve (amplitude, phase and zero-point offset) that best fit the out-of-eclipse variability at each trial period. The peak at ≈2.2 days is clear which is nearly identical to the orbital period of the binary, thus indicating the stars are tidally locked. Fig. 5. -- The rectified light curves for the Aa (left) and Ab (right) eclipses of BD -- 22◦5866 with the best model fit. -- 21 -- ) s / m k ( V R 120 100 80 60 40 20 0 -20 -40 -60 -80 -100 -120 0 0.1 0.2 0.3 0.4 0.5 Orbital Phase 0.6 0.7 0.8 0.9 1 Fig. 6. -- Five radial velocities measured for Aa (circles) and Ab (squares) of BD -- 22◦5866. The best fit sine curves assume e = 0 and equal velocity amplitudes and is fit only to the Mauna Kea data (white points) outside of eclipse. The La Silla data are plotted as black points and display a significant shift in γ due to the orbital motion of the two binaries. Measurement errors in both RV and phase are within the size of the points. The fits give KAa = KAb = 86.3 km s−1 and γ = −5.1 km s−1. -- 22 -- Fig. 7. -- Mass versus radius for known main sequence eclipsing binaries with masses, M< 0.7 M⊙. The new M-dwarf EB components, BD -- 22◦5866 Aa and Ab are shown as solid circles. The literature data on known M dwarf-M dwarf EBs (open circles) are taken from the references listed in Table 3. The lines show the theoretical, solar metallicity mass-radius relation with ages of 10 Myr (dot-dashed), 50 Myr (dashed), 500 Myr (dotted), and 5 Gyr (solid) (Baraffe et al. 1998). -- 23 -- Fig. 8. -- The maximal orbit of the low-mass quadruple-lined SB, BD -- 22◦5866 drawn nearly to scale.
astro-ph/0111459
1
0111
2001-11-23T15:13:35
Multi-wavelength spectrophotometry of EX Hydrae
[ "astro-ph" ]
We present phase-resolved infrared and optical spectrophotometry of the intermediate polar EX Hya supplemented by archival ultraviolet data. The spin-modulated emission from the accretion funnel and the emission from the accretion disk or ring contain substantial optically thin components. The white dwarf dominates the unmodulated flux in the ultraviolet and is identified by numerous absorption lines. Metal absorption in the accretion curtain may add to the observed spectral features. The secondary star is of spectral type M4+-1 and is detected by its ellipsoidal modulation. We derive a distance of 65+-11 pc which makes EX Hydrae one of the closest cataclysmic variables with a known distance. The luminosity derived from the integrated overall spectral energy distribution is 3x10^32 erg/s. The accretion rate of 3x10^15 g/s (for an 0.6 Msun white dwarf) is in reasonable agreement with the rates expected from angular momentum loss by gravitational radiation and from the observed spin-up of the white dwarf
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. (will be inserted by hand later) 1 0 0 2 v o N 3 2 1 v 9 5 4 1 1 1 0 / h p - o r t s a : v i X r a Multi-wavelength spectrophotometry of EX Hydrae ⋆ S. Eisenbart, K. Beuermann, K. Reinsch, and B.T. Gansicke Universitats-Sternwarte, Geismarlandstr. 11, D-37083 Gottingen, Germany Received October 1, 2001 / Accepted November 23, 2001 Abstract. We present phase-resolved infrared and optical spectrophotometry of the intermediate polar EX Hya supplemented by archival ultraviolet data. The spin-modulated emission from the accretion funnel and the emission from the accretion disk or ring contain substantial optically thin components. The white dwarf dominates the unmodulated flux in the ultraviolet and is identified by numerous absorption lines. Metal absorption in the accretion curtain may add to the observed spectral features. The secondary star is of spectral type M4±1 and is detected by its ellipsoidal modulation. We derive a distance of 65 ± 11 pc which makes EX Hydrae one of the closest cataclysmic variables with a known distance. The luminosity derived from the integrated overall spectral energy distribution is 3 × 1032 erg s−1. The accretion rate of 3 × 1015 g s−1 (for an 0.6 M⊙ white dwarf) is in reasonable agreement with the rates expected from angular momentum loss by gravitational radiation and from the observed spin-up of the white dwarf. Key words. Stars: individual: EX Hydrae -- cataclysmic variables -- intermediate polars -- accretion 1. Introduction EX Hya was discovered by Kraft (1962) and quickly recognised as an eclipsing system with an orbital pe- riod Porb = 98 min. A second prominent periodicity Pspin = 67 min was interpreted as the rotation period of the white dwarf (Vogt et al. 1980, Kruszewski et al. 1981) which led to the intermediate polar model of EX Hya (Warner 1983). X-ray emission in the quiescent state was discovered by Watson et al. (1978), a hard X-ray eclipse was first seen by Beuermann & Osborne (1985, 1988). In this paper, we report new phase-resolved infrared and optical spectrophotometry. We have identified the sec- ondary star in the infrared and use its K-band flux to de- termine the distance to EX Hya, a key parameter for the discussion of its luminosity and accretion rate. The derived distance is smaller than thought previously, requiring a current accretion rate of only 3 × 1015 g s−1 (for an 0.6 M⊙ white dwarf), consistent with the rates expected from gravitational radiation and implied by the observed spin- up of the white dwarf (Hellier & Sproats 1992). Combining Send offprint requests to: [email protected] ⋆ Based on observations made at the European Southern Observatory, Chile, with the ESO/MPI 2.2-m telescope in MPI time and at the Cerro Tololo Interamerican Observatory, Chile, under ESO programme ID 59.D-0724, and on observa- tions made with the NASA/ESA Hubble Space Telescope, ob- tained from the Data Archive at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. our data with archival IUE and HST spectrophotometry, we determine the contributions to the overall flux dis- tribution from the rotating funnel, the accretion disk or ring, and the white dwarf. The short-wavelength ultravi- olet spectrum contains metal absorption lines which are a clear signature of the white dwarf and/or absorption in the accretion funnel which surrounds the white dwarf. Based on these findings, we discuss the overall energy bal- ance and present an internally consistent description of EX Hya as an intermediate polar. 2. Observations and Data Analysis 2.1. Observations and Archival Data We have observed EX Hya in February 1997 during three subsequent nights in the near infrared (0.8 -- 2.4 µm) with the Infrared Spectrometer (IRS) at the 4-m telescope of the Cerro Tololo Interamerican Observatory (CTIO) and during one night in the optical (3600 -- 10200A) with the ESO Faint Object Spectrograph and Camera 2 (EFOSC 2) at the ESO/MPI 2.2m telescope at La Silla. We com- plemented these data by archival time-resolved ultravi- olet spectrophotometry taken with HST and IUE in 1995. Table 1 summarizes the relevant information. During all observations EX Hya was in a quiescent state. For com- parison and calibration purposes in the near infrared and optical ranges, spectra of several M dwarfs with spectral types between dM2 and dM5.5 and of an A0 standard star were taken. 2 S. Eisenbart et al.: Intermediate Polar EX Hydrae Table 1. Summary of time-resolved observations used for the present analysis. tint and tres are the exposure times and and the effective time resolution. The values quoted for the HST FOS data refer to time-averaged spectra (see text). Telescope Instrument λ (A) date / UT CTIO 4-m IRS 8200 -- 24200 Feb 22/23 1997 05:55 -- 09:55 Feb 23/24 1997 07:35 -- 09:20 Feb 24/25 1997 07:00 -- 09:55 tint (sec) tres Observer (sec) 60 65 K. Reinsch S. Eisenbart ESO 2.2-m EFOSC2 3600 -- 10200 Feb 28/Mar 1 1997 06:50 -- 08:50 240 ∼ 360 K. Reinsch S. Eisenbart IUE HST SWP FOS 1150 -- 1980 Jun 23 1995, 13:50 -- Jun 24 1995, 20:30 600 ∼ 2400 K. Mukai 1160 -- 1600 Feb 12 1995 04:05 -- 14:20 Feb 15 1995 12:30 -- 22:40 3.34 ∼ 13.3 S. R. Rosen 2.1.1. IRS 2.1.3. IUE The IRS at CTIO is a cooled grating spectrometer equipped with a two-dimensional InSb detector array which allows spectral resolutions of λ/∆λ ≈ 150...3000 in the wavelength range 0.8 − 5 µm (Depoy et al. 1990). When used by us, it was upgraded compared to Depoy et al. We chose a setup which allowed to observe the whole wavelength range of 8300 − 24200 A with a spectral reso- lution of λ/∆λ ≃ 560 and a time resolution of 65 s. The seeing was 1.0′′ -- 1.3′′ during the first and 1.2′′ -- 1.8′′ during the second and third night. A fraction of the observation time in nights 2 and 3 was affected by clouds and the spec- tra taken during these times were discarded. All spectra were obtained through a 2′′ aperture. A total of 222 useful spectra of EX Hya were taken, the majority during the first night. In order to correct the spectra for water vapor absorption in the infrared, we observed the nearby G5 star SAO 181198 at regular intervals. The NIR spectrum of this star has virtually no intrinsic spectral features apart from the hydrogen lines and the hourly spectra taken of this star allow a correction for H2O absorption, at least on this time scale. The M dwarfs observed were Gl 382 (dM2), Gl 443 (dM3), Gl 402 (dM4), Gl 285 (dM4.5e), and Gl 473AB (dM5.5e). As for EX Hya, spectra of nearby G0 -- G5 stars were taken for correction purposes. 2.1.2. EFOSC 2 The optical spectra were taken a few days after the in- frared observations when EX Hya was practically at the same brightness level. The sky was clear with a seeing of 1.7′′. We observed EX Hya through a 2′′ slit with low spec- tral resolution (EFOSC2 grating G1, λ/∆λ ≃ 150) cov- ering the wide wavelength range from 3600 to 10200 A. A total of 20 spectra were taken with an integration time of 4 min and an effective time resolution of 6 min. They cover 1.2 orbital or 1.7 spin cycles and include two eclipses. From the IUE data archive, we selected a continuous set of 45 SWP spectra (λ = 1150−1980 A) obtained by K. Mukai over 1.3 days in June 1995. They are the same spectra used by Mauche (1999) in his analysis of EX Hya. The spectra were taken through the large aperture with an exposure time of 600 s (∆φ67 = 0.15). For further description of the data including a list of the phases of the individual spectra see Mauche (1999). The SWP spectra were supplemented by long-wavelength spectra which are effectively not spin- resolved, however. 2.1.4. HST From the HST archive, we obtained two sets of spectra taken by S. R. Rosen with the Faint Object Spectrograph (FOS) on Feb 12 and Feb 15, 1995. The spectra were ob- tained with the G130H grating and the 4.3′′ aperture cov- ering the wavelength range 1160 -- 1600 A at a FWHM res- olution of 1.5 A. There are 14 data sets of 120 spectra each which extend over orbital phases φ98 = −0.15 to +0.20 and thereby cover 14 eclipses. The data were taken with the FOS operated in rapid readout mode at an integra- tion time of 3.34 s, resulting in an effective time resolution of 13.3 s. In order to improve the S/N ratio of the in- dividual spectra, we averaged 12 subsequent spectra and thus reduced each of the 14 data sets to 10 spectra with a S/N ≃ 8 and an effective time resolution of 160 s. We note that the eclipse in EX Hya of about 3 min full width is seen in the original data, but is lost at the reduced time resolution. 2.2. Data analysis The IUE and HST data required no further reduction, except for the removal of small spectral sections in the IUE data which were affected by reseau marks. The optical spectra were processed with standard MIDAS routines. S. Eisenbart et al.: Intermediate Polar EX Hydrae 3 The data structure and the intense sky background of the infrared spectra required special routines which one of us (SE) developed and integrated into MIDAS. Absolute flux calibration of near IR spectra is diffi- cult because suitable spectrophotometric standard stars are rare. To overcome this problem, we took spectra of several A-type photometric standard stars. The near IR spectra of such stars are blackbody-like and almost fea- tureless except for H and He absorption lines. This al- lows us to construct the intrinsic spectrum of each star as a blackbody for the effective temperature of its spectral type and to reproduce observed photometric JHK mag- nitudes by appropriate least-squares scaling. In all cases, the colours of the constructed spectra agreed with the lit- erature values to better than ±0.05 mag. Finally, the in- dividual spectral response functions derived from the con- structed and the observed spectra agreed with each other to within a few percent over the whole wavelength range. Using this procedure, the mean atmospheric absorption in the object spectra is already accounted for in the spectral response function. The remainder was removed using the spectra of the 'monitor' stars mentioned in section 2.1.1. 3. Light curve analysis EX Hya shows variations on the 67-min spin period and on the 98-min orbital period. For the light curve analy- sis, we used the orbital and spin ephemerides of Hellier & Sproats (1992), including the sinusoidal term in the former. Both ephemerides are given in Julian Days and refer to the barycentric dynamical time definition. Phase φ98 = 0 indicates the center of the optical (and X-ray) eclipse and phase φ67 = 0 the maximum of the optical (and X-ray) flux during the spin cycle. We confirm the finding by other authors (Hurwitz 1997, Mauche 1999; Mukai et al. 1998) that the orbital ephemeris needs an update (the eclipse appears at φ98 = −0.02 instead of at 0.00. This small deviation is not of relevance for the present study. The flux-calibrated time-resolved spectra allow an analysis of the spectral variations as a function of phase and a light curve analysis for finite bands in wavelength. We start here with the light curve analysis. From the individual spectra, we extract the time de- pendence of the total fluxes in broad bands (dominated by the continuum) and of the fluxes in the prominent emission lines above the continuum. The optical and IR broad bands approximately agree with Bessell UBVRI and Johnson IJHK, respectively, while the ultraviolet bands were, in part, chosen to be free of emission lines. We fit the individual data trains of the continuum and the emis- sion line fluxes with model light curves of the form F (t) = C + B t + A67 cos 2π (φ67(t) + ∆φ67) +A98 cos 2π (φ98(t) + ∆φ98) −A49 cos 2π (φ49(t) + ∆φ49) (1) where the term B t accounts for a slow variation of the mean flux over the observation time, A67, A98, and A49 are the amplitudes of the modulations at the spin, the or- bital, and half the orbital period, φ67, φ98, and φ49 are the corresponding phases, and ∆φ67, ∆φ98, and ∆φ49 the phase shifts relative to the ephemerides of Hellier & Spoats (1992). (Note that a phase shift ∆φ49 = 0.02 corresponds to a shift by 0.01 orbital periods). The only significant modulation in the ultraviolet occurs at the spin period. Spin modulation is also dominant in the optical and our data train is too short to obtain meaningful parameters for modulation at the other periods. Only the longer time cov- erage in the infrared allows to deduce the modulation at all three periods provided for in Eq. 1. Tables 2 and 3 list the least-squares fit parameters. Of the emission lines, Pγ is covered, too, but can not sufficiently well be separated from Heiλ10830. The mean visual and K-band fluxes in Table 2 correspond to mean magnitudes of EX Hya of V = 13.43 and K = 11.77 at the time of our observations. 3.1. Spin modulation Figure 1 (upper panels) show examples of the IUE fluxes fitted with line 1 of Eq. (1): (i) the total flux in the 1250 -- 1950 A range and (ii) the emission line flux of Civλ1550. Both fits benefit from including the term B t in Eq. (1). We conclude that there is a gradual increase of the ultra- violet flux over the time of the observation on top of the spin modulation. The folded data, after subtraction of the linear term, A + B t, are shown in the lower panels. Tables 2 and 3 list the mean spectral fluxes f , the integrated line fluxes F , the corresponding amplitudes A67, phase shifts ∆φ67, and the corresponding fractional modulations, e.g., M67 = 2 A67/fmax, where fmax is the mean spectral flux at spin maximum. M67 = 1 corresponds to 100 % mod- ulation and M67 > 1 for Siiiiλ1300 implies that the line which is in emission at spin maximum turns into absorp- tion at spin minimum. Allowance for a time-dependent amplitude would further improve the fit in Fig. 1, but com- plicate the definition of the modulation amplitude. The HST data yield very similar spin modulation but do not require the term B t. The modulation of the line fluxes exceeds the modulation of the continuum in the HST and IUE data as well as in the optical data (Tables 2 and 3). The infrared data continue the decline in the amplitude of the spin modulation seen in the optical bands. Fig. 2 shows the result for the K-band (left top panel) and the Pβ line (right top panel). Optical depth effects in the in- frared emission line region are important as indicated by Bγ. This line displays practically zero spin modulation of the integrated line flux (Table 2), but varies in width such that broad emission lines with deep absorption cores ap- pear in the spin-modulated spectral flux (spin maximum -- spin minimum). Fig. 3 (bottom panel) shows Bγ and the absorption cores of the higher Brackett lines up to B13. Such effects are still present in Pβ and decrease in the higher Paschen lines. 4 S. Eisenbart et al.: Intermediate Polar EX Hydrae Fig. 1. Fits of line 1 of Eq. (1) to sample IUE light curves of EX Hya. Left panel: total flux in the 1250 -- 1950 A band, right panel: Civλ1550 emission line flux. Top panels: Original time series and fitted model, bottom panels: data with the linear term A + B t subtracted and folded over the 67-min spin phase. Fig. 2. Left panels: Spin and orbital modulation of EX Hya in the IR (quasi K−band: 20000 − 24000 A). The data points around the eclipse (grey) were excluded from the fits. Right panels: Same for the Paschen Pβ line. The fit in the lower right panel represents the sum of the modulations at the orbital and half the orbital period. 3.2. Orbital modulation In our optical spectrophotometry, the orbital eclipse oc- curs slightly early at φ98 = 0.98. There is an indication for a shallow maximum near φ98 ≃ 0.8 as shown to be present by Hellier et al. (2000, see also Siegel at al. 1989). The light curves of the Balmer line fluxes are flat and show a weak broad eclipse, as expected if part of the emission line fluxes originates from the accretion disk or ring. Our infrared continuum data (lower left panel of Fig. 2) show the orbital eclipse superposed on substantial modu- lation at half the orbital period (data points covering the eclipse are displayed in gray). The infrared eclipse is more pronounced at spin maximum than at spin minimum, but is present at all spin phases. At the same time, the infrared eclipse is much wider than the eclipses at all other wave- lengths, indicating that a much more extended structure than just the accretion funnel is partially eclipsed. The full width of about 10 min corresponds to an eclipsed object of S. Eisenbart et al.: Intermediate Polar EX Hydrae 5 Table 2. Mean spectral fluxes f in box-car shaped wave- length intervals and amplitudes of modulation at the spin A67 in units of 10−16 erg cm−2s−1A−1. ∆φ67 is the shift in the spin maximum relative to the ephemeris of Hellier & Spoats (1992) and M67 the modulation amplitude defined in the text. The lower section gives the corresponding in- formation for the modulation at half the orbital period in the infrared. Table 3. Mean integrated line fluxes F and modulation amplitudes A67 in units of 10−16 erg cm−2s−1. ∆φ67 and M67 are as in Table 2. The lower sections give the cor- responding information for the modulation at the orbital and half the orbital period for the Cerro Tololo infrared data. Emission line F A67 ∆φ67 M67 λλ (A) f A67 ∆φ67 M67 a) Hubble Space Telescope: a) Hubble Space Telescope: 1255 -- 1285 1350 -- 1365 1350 -- 1380 1425 -- 1525 1570 -- 1600 1480 1590 1630 1500 1360 319 ± 24 276 ± 22 304 ± 23 221 ± 19 261 ± 17 −0.010 +0.01 +0.005 +0.005 +0.015 b) International Ultraviolet Explorer: 1255 -- 1285 1350 -- 1380 1425 -- 1525 1575 -- 1625 1680 -- 1780 1875 -- 1975 1450 1560 1440 1220 1030 1030 c) ESO EFOSC2: 264 ± 40 356 ± 60 211 ± 52 250 ± 49 262 ± 35 283 ± 39 3600 -- 4000 3900 -- 5000 5000 -- 6200 5800 -- 8000 7500 -- 10000 539 286 155 121 75 140 ± 18 59 ± 11 26 ± 6 22 ± 4 11.4 ± 2.4 d) Cerro Tololo IRS: +0.05 +0.015 +0.02 +0.035 +0.025 +0.025 +0.03 +0.035 +0.05 +0.04 +0.04 9650 -- 10650 11000 -- 13500 14500 -- 18000 20000 -- 24000 52.0 37.6 18.5 8.3 5.28 ± 0.32 +0.035 3.31 ± 0.22 +0.03 1.34 ± 0.10 +0.03 0.47 ± 0.05 +0.025 0.355 0.296 0.314 0.256 0.323 0.308 0.372 0.256 0.341 0.407 0.430 0.411 0.342 0.287 0.312 0.265 0.184 0.162 0.135 0.107 λλ (A) f A49 ∆φ49 M49 9650 -- 10650 11000 -- 13500 14500 -- 18000 20000 -- 24000 52.0 37.6 18.5 8.3 2.39 ± 0.34 −0.005 2.16 ± 0.23 −0.02 1.21 ± 0.11 −0.03 0.56 ± 0.05 −0.04 0.088 0.109 0.123 0.127 C III 1175 Ly α N V 1240 Si III 1300 C II 1335 Si IV 1400 C IV 1550 13000 27300 10300 6190 9400 23300 56400 11900 ± 810 +0.045 13000 ± 1000 +0.035 +0.005 6990 ± 480 +0.04 7720 ± 260 +0.035 4870 ± 390 17300 ± 660 +0.02 37500 ± 1400 +0.00 b) International Ultraviolet Explorer: Si IV 1400 C IV 1550 He II 1640 Al III 1860 21300 51600 6740 4960 14100 ± 1800 +0.015 39400 ± 3000 +0.005 +0.045 +0.030 8490 ± 980 4560 ± 720 c) ESO EFOSC2: He II 4686 H β He I 5875 H α 1413 11288 2288 10113 453 ± 58 5135 ± 350 1114 ± 73 3463 ± 238 +0.025 +0.01 −0.005 +0.01 d) Cerro Tololo IRS: 0.954 0.646 0.807 1.11 0.682 0.853 0.798 0.796 0.865 1.11 0.958 0.484 0.625 0.654 0.510 P δ P β B γ 2060 2080 585 412 ± 14 334 ± 15 8 ± 5 +0.020 +0.020 +0.000 0.333 0.277 0.026 Emission line F A98 ∆φ98 M98 P δ P β B γ 2060 2080 585 139 ± 17 157 ± 17 40 ± 8 +0.330 +0.275 +0.240 0.126 0.140 0.130 Emission line F A49 ∆φ49 M49 P δ P β B γ 2060 2080 585 48 ± 15 100 ± 15 30 ± 5 +0.025 +0.015 −0.025 0.045 0.092 0.098 about 3 × 1010 cm diameter centered on the location of the white dwarf, quite likely the accretion disk or ring filling about 80% of the Roche lobe. A shallow wide eclipse was also seen in the optical data of Siegel et al. (1989) in ad- dition to the narrow eclipse of the accretion funnel. In our infrared data, the time resolution is not sufficient to dis- entangle the funnel and disk contributions to the eclipse. The central depth of the mean K-band eclipse is about 1.2 × 10−16 erg cm−2s−1A−1 or 1.8 mJy. The infrared emission lines display a different orbital modulation and a weaker eclipse. Fig. 2 (lower right panel) shows P β as an example. The light curve is practically flat in the interval φ98 = 0.5 − 1.0 and shows a hump at φ98 = 0.25 which is phase shifted with respect to the blue continuum hump (Hellier et al. 2000) by roughly 180◦. The fit shown along with the data includes the 98-min and the 49-min variations which represent the first two terms of a Fourier expansion P Ancos 2π n φ98 of the orbital light curve. We interpret the maximum of the Paschen line flux as due to the inner illuminated side of the bulge, whereas the blue continuum originates from its outside heated by the interaction of the stream with the disk. This bulge is also detected by its ultraviolet line absorption (see Sect. 4.4 below). 6 S. Eisenbart et al.: Intermediate Polar EX Hydrae 3.3. Ellipsoidal modulation of the secondary star In the infrared, the principal orbital modulation of the continuum is a double wave of 49 min period. There is no significant modulation at the orbital period. The mini- mum at φ98 = 0.98 coincides with the center of the eclipse to within 0.01 in φ98. Fig. 2 (left lower panel) shows the variation of the K-band flux, with the other infrared bands behaving similarly. Table 2 lists the relevant fit parame- ters for our four infrared bands. The modulation depth is M = 0.13 in K and decreases to 0.09 in the quasi-I band. The double wave looks like the ellipsoidal modulation of the secondary star. While such a modulation is ex- pected, studies of other CVs show that other light sources can mimic the ellipsoidal modulation. In the dwarf nova IP Peg, part of the double hump structure likely occurs in the accretion disk (Froning et al. 1999) and in the magnetic system AR UMa, beamed cyclotron emission may con- tribute (Howell et al. 2001). Model calculations of Kube et al. (2000) for magnetic CVs demonstrate that a double hump can also result from the accretion stream. Typically, however, the energy distributions of the additional light sources differ from that of the secondary star and phase- resolved spectrophotometry should be able to disentangle their contributions. E.g., in low-state spectrophotometry of the magnetic system UZ For, the ellipsoidal modula- tion in the near infrared can safely bex separated from the bluer cyclotron source (Schwope et al. 1990). Considering these examples, we accept that some of the observed double hump modulation in EX Hya may originate from the accretion disk, but assume also that this contribution is less important than in other CVs as, e.g. IP Peg. The reason is that EX Hya displays no equiv- alent to the double hump in the optical continuum (Siegel et al. 1989, Hellier et al. 2000) nor in the optical and in- frared emission lines. Also, the accretion disk in EX Hya is not well developed (King & Wynn 1999). Another ef- fect which could affect the observed double humped light curve is the eclipse of the secondary star at superior con- junction (φ98 ≃ 0.48) by the accretion disk. If the disk is largely optically thin, however, as suggested below, the magnitude of this obscuration would be small. In order to account for any other contribution to the 49-min in- frared modulation in EX Hya, we take the amplitude of the ellipsoidal modulation to be Aell = α A49 with α ≤ 1, where Aell is the amplitude of the ellipsoidal modulation at φ98 = 0.5. In estimating α, we refer to the analysis of the ellipsoidal modulation in IP Peg by Froning et al. (1999) which yields α ≃ 0.5. For the reasons given above, we expect that in EX Hya a larger fraction is of ellipsoidal origin and adopt α = 0.6 . . . 1.0 as a plausible range. We determine the flux f of the secondary star from Aell and an expression for the relative modulation Aell/f . To this end, we again refer to the numerical study of Froning et al. (1999) who carefully determined the modulation using the H-band gravity and limb darkening coefficients of a Roche lobe filling secondary star of Teff = 3000 K, similar to that in EX Hya. Their result can be scaled as Aell,H/fH = 0.13 sin2i 1 + q , (2) where fH is the equivalent spectral flux of a spherical star, q = M2/M1 is the mass ratio, i is the orbital inclination, the functional form of Eq. 2 is that of the dominant term in Kopal's analytic expression (Binnendijk 1974), and the index H refers to the H-band. With i = 79◦ (Hellier et al. 2000), q ≃ 0.19 ± 0.04 (Hellier et al. 1987, Vande Putte et al. 2001), and α = 0.8 ± 0.2 we then have fH = 7.7 αA49,H 1 + q sin2i = (7.6 ± 2.0) A49,H (3) where the error accounts for the uncertainties in α, A49,H, and q. With A49,H from Tab. 2, we obtain the H-band flux of the secondary star in EX Hya as fH = (9.2 ± 2.4) × 10−16 erg cm−2s−1A−1 or 8.0 ± 2.1 mJy, where most of the error is systematic in nature and represents the full ac- ceptable range, not a standard deviation. Using the nu- merical factor of Eq. 2 for all infrared bands, we obtain the fluxes plotted in Fig. 4 (open circles, statistical errors only). Because of the wavelength dependence of limb dark- ening (e.g. Al-Naimiy 1978), the true spectral energy dis- tribution may be slightly redder (with a pivot point at H). This is consistent with the spectral distribution of the secondary star determined below (Sect. 4.3). Of the total orbital mean K-band flux of 13.3 mJy, this interpretation then assigns 7.3 mJy to the secondary star. Its infrared magnitude is K = 12.43+0.24 −0.30. We continue the discussion of the spectrum of the secondary star in Sect. 4.3. Of the remaining mean K-band flux, we attribute 1.2 mJy to the the accretion funnel and 4.8 mJy to the accretion disk and white dwarf (Sects. 4.2 and 4.4). Since about 1/2 of the funnel emission and about 1/5 of the disk are eclipsed, we expect a central depth of the partial eclipse of about 1.7 mJy or 1.1 × 10−16 erg cm−2s−1A−1, equal to the observed depth within the uncertainties (Sect. 3.2). 4. Spectral analysis 4.1. Overall spectrum Figure 3 (upper left panel) shows the mean overall spectral energy distribution of EX Hya from the ultraviolet to the infrared. The mean IUE and HST spectra agree in abso- lute flux (Table 2) and, in the region of overlap, we show the less noisy HST spectrum. We have also schematically included the ORFEUS ii continuum of Mauche (1999) which agrees reasonably well with the HUT spectrum of Greeley et al. (1997). In judging the overall spectrum, one should note that the individual sections were not ob- served contemporaneously. In spite of this, an astound- ingly coherent view emerges. The spectrum clearly shows the Balmer, the Paschen, and the Brackett jumps in emis- sion and, hence, contains substantial optically thin com- ponents, as suspected already by Berriman et al. (1985) on the basis of the infrared colours of EX Hya. S. Eisenbart et al.: Intermediate Polar EX Hydrae 7 Fig. 3. Top: Mean spec- trum of EX Hya from 1000 to 24000 A. Bottom: Difference between spin maximum and spin mini- mum spectra. See text for the long wavelength IUE part of the spectrum (dot- ted). 4.2. Spectrum of the spin-modulated component The spin-modulated component has been extracted at all wavelengths and the amplitudes A67 of the broad bands and the line fluxes are listed in Tables 2 and 3. The dif- ference between the spectral fluxes at spin maximum and spin minimum, 2 A67(λ), is depicted in the lower panel of Fig. 3. In the long-wavelength IUE range, the modulation amplitude could not be measured because of the lack of time resolution. We have, therefore, represented this spec- tral section schematically by the dotted line which has the mean slope of the long-wavelength spectrum and is ad- justed to the modulated flux of the short wavelength IUE spectrum. The modulation of the ORFEUS ii spectrum at 1010 A (Mauche 1999) is indicated by a single data point. The spin-modulated component displays emission lines of the Balmer and Paschen series and the associ- ated jumps in emission, while the Brackett lines appear in absorption (see above). The mean full base width of the ultraviolet emission lines corresponds to velocities of ±2300 km s−1(see Fig. 8, below), typical of the funnel at a couple of white dwarf radii. Spin modulation originates from the rotating partially optically thick accretion funnels at both poles of the white dwarf. Because any optically thin and non-occulted com- ponent is constant in time, the mean funnel flux ffun(λ) ≥ A67(λ). We need, therefore, an additional assumption to deduce ffun(λ) from A67(λ). We note that the bright ul- traviolet lines of Ciii, Civ, Siiii, Siiv, and Nv are strongly modulated (see M67 in Tab. 3). For definiteness, we refer to the 80% modulation of the strongest ultraviolet line, Civλ1550, and assume that this line originates entirely in the funnel which is not unreasonable because, as shown be- low, Civ absorption is weak in the white dwarf spectrum. Adopting the modulation of Civλ1550 as representative of ffun(λ), we compute the mean funnel flux to a first-order approximation as ffun(λ) = 1.5 A67(λ). (4) In the optical continuum, the eclipsed flux at spin maxi- mum slightly exceeds the spin modulated flux which sug- gests a similar factor. Nevertheless, there are clearly un- certainties in defining ffun(λ) by a simple relation as in Eq. 4. We expect the ultraviolet emission of the heated polar caps of the white dwarf to display a spin modula- tion, too, and separation of the two spin-modulated com- ponents becomes non-trivial (see Sect. 4.5). On the other hand, ffun(λ) clearly becomes small at long wavelengths and the implied error in the unmodulated component be- comes then small, too. We use Eq. 4 to compute the con- 8 S. Eisenbart et al.: Intermediate Polar EX Hydrae Fig. 4. Contribution of the secondary to the mean spectrum of EX Hya. The sample M-star spectra of spectral types M3.0, M4.5 and M5.5, respectively, have been scaled to the expected fluxes (crosses) derived from the ellipsoidal modulations listed in Table 2 (see text). For clarity, the noisy region between the J and H-bands has been smoothed. tribution of the funnel emission to the luminosity, but do not attempt to model its spectral distribution. 4.3. Spectrum of the secondary star Fig. 4 shows the quasi-Johnson IJHK fluxes of the sec- ondary star as defined above (Sect. 3.2.2.) along with the spectra of M3 to M5.5 stars least squares fitted to the data points for α = 0.8. The systematic uncertainty in the flux is at most 2 mJy (Sect. 3.3). Formally, the four data points are best fitted by an M3 star, but, as noted above, the true spectrum will be slightly redder and the spectral type later, about M3 to M5, still consistent with the M3 assignment by Dhillon et al. (1997). The contribution of the secondary star to the total spectrum is largest at long wavelengths and the only pronounced distinctive feature in this range is the broad hump which is centered at ∼ 2.2 µm and flanked by H2O absorption dips. This feature is clearly visible also in the spectrum of EX Hya. We determine the most probable contribution from the secondary star by sub- tracting the spectrum of an M3, an M4.5, and an M5.5 star from the observed spectrum. Fig. 5 shows the corre- sponding difference spectra. We expect this spectrum to decrease smoothly since the main contributions should be from the spin-modulated component (Fig. 3) and the ac- cretion disk (Fig. 7) and both decrease with wavelength. The smoothest decrease is obtained for the M4.5 star, in agreement with the expected spectral type in a CV with 98 min orbital period (Beuermann et al. 1998). We adopt a spectral type M4 ± 1. The detection of absorption lines from the secondary in EX Hya is difficult. Fig. 6 compares the spectrum of the M4.5 dwarf Gl 285 with spectra of EX Hya collected near spin minimum: (i) mean spin minimum, (ii) ellip- soidal maximum (φ98 ≃ 0.75) near spin minimum, and (iii) eclipse (φ98 ≃ 0, backside of star) near spin minimum. The M-star features are largely obscured by lthe emission line components in the 8300 -- 9600 A range. In the K-band, the presence of the M-star is indicated by the faint NaI doublet and the CO molecular bands( see also Dhillon et al. 1997). Although our spectral resolution is not sufficient to resolve the NaI lines, phase-shifted superposition of the spectra suggests that NaI absorption is preferentially present on the dark side of the star and displays a radial velocity consistent with the motion of the secondary star (an un- certain 400 km s−1). Higher resolution is needed to settle the question of the radial velocity of the secondary star (Vande Putte et al. 2001). 4.4. Spectrum of the accretion disk Figure 7 shows the (slightly smoothed) spectrum which remains after subtraction of the funnel component ffun(λ) and the secondary star, assuming it has K = 12.43, and is a dM4.5 star (Figs. 4 and 5). We assign this spectrum to the sum of the accretion disk including the bright spot at its edge and the white dwarf. The spec- trum lacks the intense ultraviolet lines which belong to the funnel component, but contains a substantial fraction of the Hei, Balmer, Paschen, and Brackett line emission. Mgiiλ2800 may, in part, originate in the funnel and be falsely assigned to the disk because we subtracted the long-wavelength IUE contribution to the funnel spectrum only in a schematic way without accounting for the emis- sion lines. The (unsmoothed) short-wavelength section of the spectrum in Fig. 7 is shown enlarged in the insert and is further discussed in Sect. 4.5 (Fig. 9). This part of the spectrum is dominated by the heated white dwarf. We model the λ > 3000 A part of the spectrum by an isobaric and isothermal disk of pure hydrogen plus a blackbody representation of the white dwarf. The spectrum of the disk is computed following the description of Gansicke et al. (1997). It has an outer radius of 1.6 × 1010 cm, a thick- S. Eisenbart et al.: Intermediate Polar EX Hydrae 9 Fig. 5. Mean spectrum of EX Hya (top) with the scaled sample M-star spectra from Fig. 4 sub- tracted. The individual spectra have been offset by multiples of 0.5 flux units for clarity. Fig. 6. Spectral signatures of the secondary star in the IR spectrum of EX Hya. Both panels show the spectrum of the M4.5 star Gl285 and three spectra of EX Hya, (i) mean near spin minimum, (ii) 49-min maximum near spin minimum, and (iii) eclipse near spin minimum (from top to bottom). All spectra are normalised to unity at 9800 A (left panel) or 22500 A(right panel) and vertically offset by multiples of 0.5 flux units (left panel) and 0.3 flux units (right panel). Spectral features expected to be present in the secondary star are indicated. H2O refers to regions affected by atmospheric water vapour absorption. The strong emission lines in the EX Hya spectra are the hydrogen Paschen series and Caiiλ8498, 8542, 8662 (left panel) and Bγ (right panel). ness of 2×108 cm, a pressure of 200 dyne cm−2, an electron temperature of 7000 K, close to the canonical tempera- ture of an optically thin accretion disk (Williams 1980), and is seen at an inclination of 79◦. For definiteness, we assume that the disk has a central hole of 6 × 109 cm ra- dius (Hellier et al. 1987), but the structure could also be a ring with a larger inner radius surrounding the magne- tosphere of the white dwarf. This component is optically thin and accounts for much of the line emission and the Balmer, Paschen, and Brackett jumps, but lacks contin- 10 S. Eisenbart et al.: Intermediate Polar EX Hydrae of the the Slightly smoothed Fig. 7. spectrum of EX Hya after subtraction spin- modulated funnel component and the secondary star. A single data point at 1010 A denotes unmodulated flux of the ORFEUS spec- trum (Mauche 1999). Also shown are the model spec- trum of the accretion disk (dashed curve), a schematic representation of the white dwarf (dotted curve), and the sum (solid curve). The insert shows an expanded version of the unsmoothed ultraviolet part of the spectrum. nounced selfabsorption cores of highly excited ultravio- let lines are seen which disappear after the eclipse. These cores are produced in the bulge at the edge of the ac- cretion disk, i.e. the interaction region of the accretion stream with the disk or ring. This bulge produces also the enhanced Paschen line emission at φ98 = 0.25, the orbital maximum in the B-band at φ98 = 0.85 (Hellier et al. 2000), and the pre-eclipse dip in the EUV and soft X-ray regime (C´ordova et al. 1985, Rosen et al. 1988, Hurwitz et al. 1997, Mauche 1999). The radial velocity of the absorption lines in the difference spectrum (C in Fig. 8) is consistent with zero, as expected for material at the edge of the disk or ring which moves essentially perpendicular to the line of sight. The presence of Nv indicates an excitation tem- perature of the order of 50 000 K or higher. The metal lines are just resolved with a FWHM of 2.5 A, corresponding to about 400 km s−1. The Lyα absorption line in the differ- ence spectrum is wider with FWHM ≃ 5 A (1000 km s−1), suggesting an origin in warm material or material with a large velocity dispersion. The column density needed to produce the bulge absorption in Lyα at φ98∼ 0.9 is about 1019 H-atoms cm−2, substantially less than the equivalent column density of 1.3 × 1020 H-atoms cm−2 of cold mat- ter of solar composition needed to account for the ob- served EUV bulge absorption (Hurwitz et al. 1997). The difference suggests that hydrogen is largely ionized and the EUV bulge absorption is primarily due to helium. 4.5. Spectrum of the white dwarf The ultraviolet is sorption lines expected from a hot atmosphere, spectrum in Fig. 7 shows the typical ab- in- reproduced in Fig. 9. section of the It uum flux around 5000 A. We include, therefore, a second component which has a pressure of 103 dyne cm−2, an elec- tron temperature of 10 000 K, a weak Balmer jump, and is nearly optically thick. The hotter component covers only ∼ 20% of the area of the cool one and may represent the heated inner edge of the disk (but see Sect. 4.5). Fig. 7 shows the sum of the two disk components (short-dashed curve). Although this is not a formal fit, it reproduces the observed optical and infrared part of the spectrum reason- ably well. Metal line emission (mostly Feii, not included in the model) may substantially contribute to the observed excess in the range 4000. . . 5500 A. Lines of the Pfund se- ries (not included either) may contribute to the excess flux longward of 22 788 A. The spectrum in Fig. 7 is based on K = 12.43 for the secondary star. If the secondary were substantially brighter, the decreased infrared flux could not be fit by the optically thin spectrum anymore; if it were fainter, there could be additional disk flux over that depicted in Fig. 7 within the limits set in Sect. 3.3. The 2 mJy uncer- tainty in the K-band flux of the secondary, would allow an underlying cool component of the disk with a black- body temperature up to Tbb = 3300 K for a disk radius of 1.6 × 1010 cm. This is a significant result because the mass in the modeled disk is <∼ 6 × 1019 g, much less than the ∼ 1022 g needed to feed one of the rare outbursts. If the outbursts result from a disk instability, a layered disk is needed in which the component of high surface density must be cooler than the limit given above. This result is very similar to the conclusions of Gansicke et al. (1997) for the case of the SU UMa dwarf nova EK Tra. Fig. 8 shows the HST spectrum before and after eclipse, at Porb = ±(0.05 − 0.15). Before eclipse pro- S. Eisenbart et al.: Intermediate Polar EX Hydrae 11 Fig. 8. Spin averaged HST spectra of EX Hya before and after eclipse (φ98= -0.15...-0.05 and φ98= +0.05. . . +0.15, upper two spectra) and difference spectrum (bottom). The identifications of the narrow absorption lines are indicated (see text). Fig. 9. HST spectrum of EX Hya with the funnel compo- nent removed. The extension for λ > 1590 A is from the IUE spectrum. A 25 000 K white dwarf spectrum with so- lar composition is shown for comparison (offset by -- 0.1 units). cluding Ciiiλ1175, Siiiλ1260, 1265, Siii,iiiλ1294 . . . 1306, Ciiλ1335, Siivλ1394, 1403, and Siiiλ1527, 1533. Some of the lines are affected by the subtraction proce- dure by which this spectrum was created, Civλ1550, Siivλ1394, 1403, and probably Ciiiλ1175. In order to es- timate the temperature of the white dwarf (or its heated pole cap) in EX Hya, we have computed a set of white dwarf spectra covering Teff = 15 000 − 50 000 K in steps of 5000 K, using the codes TLUSTY 195 and SYNSPEC 45 (Hubeny 1988; Hubeny & Lanz 1995), assuming solar abundances and log g = 8.0 (i.e. M⊙ ≃ 0.6). The ob- served line ratios Siiii/Siii and Ciii/Cii constrain the temperature of the white dwarf photosphere to Teff = 25 000 ± 3000 K and a white dwarf model of this temper- ature qualitatively fits also the absorption lines of other heavy elements. For an distance of 65 pc (Sect. 5.2), the observed flux in the range 1300 -- 1500 A corresponds to a white dwarf radius of 6 × 108 cm. We do not attempt a detailed fit to the white dwarf spectrum for the following reasons: (i) the irradiated white dwarf has very likely a non-uniform temperature distribu- tion; (ii) the spectrum is probably altered by attenuation in the accretion curtain; and (iii) the long wavelength part of the white dwarf spectrum is heavily veiled by the ac- cretion disk and practically unobservable. We comment on these points in turn. The funnel radiation which impinges on the white dwarf will produce extended hot polar caps. Taking ac- count of these effects would introduce additional parame- ters which are not well constrained. Furthermore, heating affects the temperature structure of the white dwarf at- mosphere and thereby alters the profiles of the absorption lines. In the polar AM Her, where the FUV emission is dominated by emission from the white dwarf, we could show that irradiation results in a significantly decreased depth of the Lyα absorption line (Gansicke et al. 1998). A similar effect could explain the shallow Lyα absorption observed in EX HYa. The mean photospheric spectrum of the white dwarf is probably altered by the fact that part of the light is atten- uated on its passage through the accretion curtain. An in- dication for this is provided by the double hump structure of the spectrum shown in the insert in Fig. 7. Such spectral shape is reminiscent of the white dwarf spectrum in OY Car which is heavily veiled by metal absorption in material referred to as "Feii curtain" (Horne et al. 1994). There is, in fact, a detailed similarity between the ultraviolet spec- trum in Fig. 7 and that of the white dwarf in OY Car (Fig. 5, bottom panel of Horne et al.). Metal absorption in the funnel produces the broad dips around 1700 A and 2500 A, and the remnant spike at 1595 A. Given that funnel emis- sion and absorption is abundant at other wavelengths, we are not surprised to find funnel absorption features mixed up with photospheric ones. The strong absorption lines in- dicated in Fig. 9 are not noticeably broadened, however, and probably of photospheric origin. If part of the ultraviolet emission is due to a white dwarf with heated polar caps, we may expect that the flux shows spin-modulation. There are spin-modulated contri- butions to the ultraviolet flux by the accretion funnel and by the heated white dwarf. The best visibility of the ac- cretion funnel is when the lower pole points towards the observer (Beuermann & Osborne 1988). At this rotation phase, φ67 = 0, the heated polar caps reach their mini- mum projected area. Hence, the spin modulated compo- nent from the heated polar caps assumes its minimum when the funnel flux reaches its maximum, i.e., the two modulations will be out of phase by 180◦. The amplitude of the spin modulation will then display a minimum near 1500 A, where the spectral flux of the white dwarf reaches a maximum, as it is observed (Fig. 3). 12 S. Eisenbart et al.: Intermediate Polar EX Hydrae Finally, there will be a contribution from the unheated fraction of the white dwarf photosphere. For an effective temperature of 10 000 -- 15 000 K and 109 cm radius, the peak flux would reach only 1 -- 3 mJy and can not be identi- fied against the other components. It might be contained, e.g., in the adopted disk flux which would then have to be reduced correspondingly (Sect. 5.4). 5. System parameters 5.1. Mass and radius of the secondary star We assume that the secondary is a main sequence star which is negligibly expanded over its zero age main se- quence radius (Baraffe et al. 1998, Kolb & Baraffe 1999, see also Beuermann et al. 1998). For stars of solar compo- sition, the Baraffe et al. mass radius relation of spherical stars may be approximated by a power law of the type R ∝ M β for limited ranges of mass. Considering that the equilibrium radius of a Roche lobe filling star of the same mass is larger by about 6% (Kolb 2001), we obtain R2/R⊙ = 0.77 (M2/M⊙)0.75 (5) for masses between 0.11 and 0.40 M⊙. The relation of Patterson (1984), R2/R⊙ = (M2/M⊙)0.88, predicts radii within 5% of this for masses between 0.09 and 0.20 M⊙. Eq. (5) and Roche geometry yield M2 = 0.12 ± 0.01 M⊙ and R2 = (0.157 ± 0.010) R⊙ = (1.09 ± 0.07) × 1010 cm, where we have allowed for some systematic error in M1. If the mass exceeds 0.13 M⊙, the secondary would have to be smaller than a main sequence star. If it is as low as 0.10 M⊙, an expansion in radius by more than 10% would be required. 5.2. Distance of EX Hydrae The detection of the secondary star in EX Hya allows to obtain an estimate of the distance d by the surface brightness method (Bailey 1981), log d = (K − SK)/5 + 1 + log (R/R⊙). (6) We use the re-calibration of the K-band surface brightness of late type stars by Beuermann (2000) which yields SK = 4.37±0.22 for a dwarf of spectral type M3 -- 5 and near-solar composition. With K = 12.19 . . . 12.73 and the stellar ra- dius of the last section, we obtain a distance d = 65±11 pc. This is much lower than previous estimates which include 105 pc (Warner 1987) and > 130 pc (Berriman et al. 1985). EX Hya may be one of the closest CVs with a known dis- tance. The largest contribution to the error in d arises from the uncertainty in the K-magnitude of the secondary (Sect. 3.3). The distance is near the upper limit of the derived range if the true ellipsoidal modulation is as small as 60% of the observed 49-min variation. Because of this uncertainty and the special importance of EX Hya among all CVs, a trigonometric parallax for this object is needed. Table 4. Energy fluxes of the spectral components of EX Hya. Component Hard X-rays Soft X-rays EUV Accretion curtain Accretion disk White dwarf Total E (keV) or λ (A) Flux (10−11 erg cm−2s−1) ≥ 1.0 keV 0.28...1.0 keV 0.067...0.28 keV 0.0136...0.067 keV λ = 912...24 000 A λ = 912...24 000 A λ = 912...24 000 A 15 3 ∼ 1 ∼ 1 : 12 10 18 60 5.3. Interstellar absorption X-ray spectral fits suggest that the X-ray source is par- tially covered by substantial column densities NH of hy- drogen (e.g. Allan et al. 1998). This is material within the binary system. The detection of EUV emission (Hurwitz et al. 1997) proves that the true interstellar column den- sity is small. Fitting the narrow Lyα-absorption profile in the φ98 = 0.05 -- 0.15 spectrum of Fig. 8 yields NH ≃ (3 ± 1) × 1018 H-atoms cm−2, consistent with the non- detectable extinction, EB−V <∼ 0.05 (Verbunt 1987). At d = 65 pc, the mean space density of neutral atomic hydro- gen along the line of sight to EX Hya (l, b = 304◦, +33◦) is nH = 0.01 . . . 0.02 cm−3. 5.4. Luminosity We estimate the luminosity as L = 4 π d2 F , where F is the orbital and spin-averaged energy flux, with the secondary star subtracted and corrected for interstellar absorption using NH = 3 × 1018 H-atoms cm−2 of cold matter. There is some uncertainty in the estimate of L because the mean flux used may deviate from the 4π-average. Table 4 lists the mean fluxes in individual energy bands shortward of Lyα and for individual physical com- ponents longward of Lyα. The mean X-ray flux is col- lected from Beuermann & Osborne (1984), C´ordova et al. (1985), Rosen et al. (1988), Rosen et al. (1991), Singh & Swank (1993) and Allan et al. (1998). The observed EUV and soft X-ray flux in the 0.067-0.280 keV range is ∼ 0.7×10−11 erg cm−2s−1 (C´ordova et al. 1985, Hurwitz et al. 1997) which after correction for interstellar absorption increases to ∼ 10−11 erg cm−2s−1. There is no informa- tion on the flux between 13.6 and 67 eV and we simply add another ∼ 10−11 erg cm−2s−1. Angle-dependent absorption and scattering in the fun- nel is probably responsible for the observed low-energy depression and the spin modulation of the EUV and X- ray spectrum (e.g. Kim & Beuermann 1995). The flux contained in the spin-modulated soft and hard X-ray components, i.e. half the difference between maximum S. Eisenbart et al.: Intermediate Polar EX Hydrae 13 and minimum fluxes for photon energies > 0.28 keV, amounts to 3 × 10−11 erg cm−2s−1. Applying the Zanstra method to the flux in the Heiiλ4686 line, F4686 = 1.65 × 10−13 erg cm−2s−1 (Patterson & Raymond 1985, this work), demonstrates that an additional flux F>54 ≃ 3 × 10−11 erg cm−2s−1 is absorbed at photon energies be- tween 0.054 and 0.28 keV. The sum of these two com- ponents is the minimum reprocessed flux expected from the accretion funnel. The funnel emission will actually be higher because additional energy is absorbed from the Lyα continuum below 54 eV and from the Balmer continuum of the heated white dwarf. Furthermore, vis- cous interaction of the accreted gas in the funnel may release additional energy. The sum of all these sources should account for the mean observed funnel emission of 12×10−11 erg cm−2s−1 which appears reasonable. The lat- ter number is the wavelength-integral of the funnel com- ponent for λ > 912 A , Ffun = 1.5 R A67 dλ (see Sect. 3.1 and Fig. 3). Subtracting the funnel component and the secondary star from the mean spectrum leaves the accretion disk and the white dwarf (Fig. 7). In Table 4, the disk is represented by the integral over the dotted spectrum in Fig. 7 and the white dwarf by the flux remaining after subtraction of the dotted spectrum. The disk component contains about 1/6 of the total flux (Table 4) a bit more than expected from gravitational energy release outside ri ≃ 6 × 109 cm and a white dwarf of radius R1 = 109 cm. Note that this result does not allow an inner disk radius ri < 6 R1, consistent with the visibility of the lower pole of the white dwarf in X-rays (Beuermann & Osborne 1988). The intrinsic flux of the white dwarf can not be sepa- rately identified and is included in the contribution quoted for either the heated white dwarf or the accretion disk (Table 4). For R1 = 109 cm and an effective tempera- ture of the white dwarf of 10 000 -- 15 000 K, it amounts to only (2 − 6) × 10−11 erg cm−2s−1. Hence, the total accretion-induced flux is slighly reduced below the sum of the fluxes in Table 4 to about 5.6 × 10−10 erg cm−2s−1. This corresponds to a luminosity of L = 2.8 × 1032 erg s−1 at d = 65 pc, or L = 6.7 × 1032 d 2 100 erg s−1, where d100 is the distance in units of 100 pc. Not included in Table 4 is the secondary star which contributes only 1.5 × 10−11 erg cm−2s−1. 5.5. Accretion rate M = L R1/GM1, The accretion rate is obtained as where M1 and R1 are mass and radius of the white M = 5 × 1015 dwarf. With L from the last section, (R1/109 cm)(M1/M⊙)−1 d 2 100 g s−1. M1 is estimated from the X-ray temperature as 0.45 -- 0.48 M⊙ (Fugimoto & Ishida 1997, Cropper et al. 1999, Ezuka & Ishida 1999, Ramsay 2000). The radial velocity amplitudes of the white dwarf, K1 ≃ 69 ± 9 km s−1 (Hellier et al. 1987), and of the secondary star, K2 ≃ 356 ± 44 km s−1 (Vande Putte et al. 2001), yield M2 = 0.095 M⊙ and M1 = 0.49 M⊙. According to our discussion in Sect. 5.1, the secondary would have to be substantially expanded over a main se- quence star if of such a low mass and we prefer the higher mass of 0.12 M⊙ because, within the errors, K1 and K2 are consistent also with M2 = 0.12 M⊙ and M1 as high as 0.7 M⊙. Given the remaining uncertainty in the masses, we quote the accretion rate in terms of a standard white dwarf of 0.6 M⊙. Using a Wood (1994) white dwarf model with thick hydrogen envelope and Teff = 104 K, the radius is then R1 = 8.9 × 108 cm and M ≃ 7.5 × 1015 d 2 100 g s−1 or M ≃ 3.1 × 1015 g s−1 at d = 65 pc. This is consistent with the accretion rate as expected from gravitational radiation as the dominant angular momentum loss mechanism. It is also consistent with the distance-independent accretion rate deduced from the observed spin-up rate of the white Mspin = 2.4×1015 (ri/1010 cm)−1/2 g s−1 dwarf in EX Hya, for M1 = 0.6 M⊙ (Ritter 1985) which is derived on the as- sumption that the accreted matter couples onto the mag- netic field of the white dwarf at the inner disk radius ri. The estimated ri ≃ 6 × 109 cm (Hellier et al. 1987) yields Mspin ≃ 3.1 × 1015 g s−1. These numbers do not change substantially if M1 ≃ 0.5 M⊙. 5.6. Magnetic field strength of the white dwarf A coarse estimate of the (di -- )polar field strength of the white dwarf of Bp ≃ 0.18 MG can be derived from ri ≃ 6 × 109 cm (Hellier et al. 1987) and M ≃ 3 × 1015 g s−1 if ri is equated to rµ ≃ 2−10/7µ4/7(GM1)−1/7 M −2/7 cm, i.e. about 1/2 of the spherical Alfv´en radius (e.g. Frank et al. 1992, their Eq. 6.12). Larger ri (Hellier et al. 2000) would lower the field estimate. Such a low field is consistent with the lack of cyclotron emission in the infrared. 6. Conclusions We have detected the dM4 ± 1 secondary star in EX Hya by its ellipsoidal modulation and spectral energy distri- bution. From its infrared magnitude K ≃ 12.5, we have derived a distance of EX Hya of 65 ± 11 pc, substantially less than previous estimates. The integrated flux from the X-ray to the infrared range averaged over spin and orbital period is 6 × 10−10 erg cm−2s−1, yielding a luminosity of about 3 × 1032 erg s−1 and an accretion rate which is con- sistent with the observed spin-up and with gravitational radiation as the dominant angular momentum loss mech- anism. Distance, luminosity, and accretion rate depend upon which fraction of the observed infrared continuum modulation at half the orbital period may be interpreted as ellipsoidal modulation of the secondary star. If this frac- tion is near unity, d is near the lower limit of 55 pc, if it less than 60%, d may exceed 75 pc. This question can best be resolved by obtaining a trigonometric parallax. The heated white dwarf is detected for the first time by strong metal absorption lines which are probably of pho- tospheric origin, but may be affected also by absorption in the accretion funnel. The equivalent width ratios of the Si ii, iii, iv and C ii, iii lines suggest a near-solar metal 14 S. Eisenbart et al.: Intermediate Polar EX Hydrae abundance in the uppermost photospheric layers and an effective temperature near 25 000 K. The radius needed to account for the emission is about 6 × 108 cm. If the white dwarf is of low mass as suggested by several stud- ies and has a radius of about 109 cm then what we see are probably the heated polar caps of the white dwarf. Intrinsically, the white dwarf may be substantially cooler than indicated by the line ratios (and perhaps display a less metal-rich surface composition). The observed "white dwarf flux" corresponds to ∼ 28% of the total accretion flux, consistent with the reprocessed fraction of the down- ward flux from the post-shock region considering both, geometry and reflection probability. There is agreement that maximum light at optical and X-ray wavelengths occurs when the upper pole points away from the observer (Hellier et al. 1987, Beuermann & Osborne 1988, Rosen et al. 1988, Siegel et al. 1989, Hellier et al. 2000). At this spin phase, the light from the heated caps is likely to reach a minimum which implies antiphased modulations of the ultraviolet light from the funnel and the white dwarf surface. The minimum in the amplitude of the spin modulation near 1500 A is consistent with this notation. More extensive ultraviolet observations can pos- sibly disentangle the contributions from both components. Of the outward directed integrated X-ray flux, the spin-modulated fraction of ∼ 20% is absorbed and re- processed in the accretion funnel. The funnel spectrum is largely the result of radiative transfer of the X-ray emis- sion from the accretion column and of the ultraviolet emis- sion from the heated white dwarf. Although it is clear that the funnel is optically thick in the emission lines and partly in the continuum, the detailed funnel geometry re- mains uncertain and we have not attempted to model the resulting spectrum (Fig. 7). Better spectral resolution in the optical and infrared is needed to study the radiative transfer in detail. The emission which is left after subtraction of the sec- ondary star, the white dwarf, and the funnel emission originates from the accretion disk or ring surrounding the white dwarf magnetosphere. The spectral flux contribu- tion of much of the remaining optical and infrared light is plausibly explained by an isothermal slab of gas of radius 1.6 × 1010 cm, thickness 2 × 108 cm, and a temperature of most of the gas near 7000 K which is optically thin plus a largely optically thick component at 10 000 K which cov- ers only 20% of the emitting area of the disk. There is a tight upper limit on the blackbody temperature of an underlying optically thick component covering the entire disk of Tbb< 3300 K. In summary, we have presented an interpretation of the overall spectral energy distribution of EX Hya which explains the individual components within the standard model of a weakly magnetic intermediate polar. Acknowledgements. We thank Dave Vande Putte and Robert Smith for making their K2-measurement of EX Hya available to us prior to publication and Ivan Hubeny for providing recent versions of TLUSTY and SYNSPEC. This work was supported in part by BMBF/DLR grants 50 OR 9610 4 and 50 OR 9903 6. References Al-Naimiy H.M. 1978, Ap&SS 53, 181 Allan A., Hellier C., Beardmore A. 1998, MNRAS 295, 167 Baraffe I., Chabrier G., Allard F., Hauschildt P. 1998, A&A 337, 403 Bailey J. 1981, MNRAS 197, 31 Beuermann K. 2000, New Astr. Rev. 44, 93 Beuermann K., Osborne J.P. 1984, Proc. X-ray Astronomy '84 (Bologna), eds. M. Oda & R. Giacconi, p. 23 Beuermann K., Osborne J.P. 1985, Space Sci. Rev. 40, 117 Beuermann K., Osborne J.P. 1988, A&A 189, 128 Beuermann K., Baraffe I., Kolb U., Weichhold M. 1998, A&A 339, 518 Berriman G., Szkody P., Capps R.W. 1985, MNRAS 217,327 Binnendijk L. 1974, Vistas in Astron. 16, 61 C´ordova F.A., Mason K.O., Kahn S.M. 1985, MNRAS 212,447 Cropper M.S., Wu K., Ramsay G., Kocabiyik A. 1999, MNRAS 306, 684 Dhillon V.S., Marsh T.R., Duck S.R., Rosen S.R. 1997, MNRAS 285, 95 Depoy D.L., Gregory B., Elias J. et al. 1990, PASP 102, 1433 Ezuka H., Ishida M. 1999, ApJS 120,277 Frank J., King A.R., Raine D. 1992, Accretion Power in Astrophysics, Cambridge Astrophys. Ser., 2nd edition Froning S.S., Robinson E.L., Welsh W.F., Wood J. 1999, ApJ 523, 399 Fugimoto R., Ishida M. 1997, ApJ 474, 774 Gansicke B.T., Beuermann K., Thomas H.-C. 1997, MNRAS 283, 388 Gansicke B.T., Hoard D.W., Beuermann K., Sion E.M., Szkody P. 1998, A&A 338, 933 Greeley B.W., Blair W.P., Long K.S., Knigge C. 1997, ApJ 488, 419 Hellier C., Mason K.O., Rosen S.R., C´ordova F.A. 1987, MNRAS 228, 463 Hellier C., Sproats L.N. 1992, IBVS 3724 Hellier C., Kemp J., Naylor T. et al. 2000, MNRAS 313, 703 Horne K., Marsh T.R., Cheng F.H., Hubeny I., Lanz Th. 1994, ApJ 426, 294 Howell S.B., Gelino D.M., Harrison T. 2001, AJ 121, 482 Hubeny, I., 1988, Comput. Phys. Comm. 52, 103 Hubeny, I., Lanz, T., 1995, ApJ 439, 875 Hurwitz M., Sirk M., Bowyer S., Yuan-Kuen K. 1997, ApJ477, 390 Kim Y., Beuermann K. 1995, A&A 298, 165 King A.R., Wynn G.A. 1999, MNRAS 310, 203 Kolb U. 2001, Talk at the Gottingen Conf., ASP Conf. Ser. in preparation Kolb U., Baraffe I. 1999, MNRAS 309, 1034 Kube J., Gansicke B.T., Beuermann K. 2000, A&A 356, 490 Kraft R.P. 1962, ApJ 135, 408 Kruszewski A., Mewe R., Heise J. et al. 1981 Space Sci. Rev. 30, 221 Mauche, C. 1999, ApJ 520, 822 Mukai K., Ishida, M., Osborne, J.P. et al. 1998, ASP Conf. Ser. 137, 554 Patterson J. 1984, ApJS 54, 443 Patterson J., Raymond J.C. 1985, ApJ 292, 550 Ramsay G. 2000, MNRAS 314, 403 Ritter H. 1985, A&A 148, 207 Rosen S.R., Mason, K.O., C´ordova, F.A. 1988, MNRAS 231, 549 Rosen S.R., Mason K.O., Mukai, K. 1991, MNRAS 249,417 S. Eisenbart et al.: Intermediate Polar EX Hydrae 15 Schwope A.D., Beuermann K., Thomas H.-C. 1990, A&A 230 120 Singh J., Swank J. 1993, MNRAS 262, 1000 Siegel N., Reinsch K., Beuermann K., van der Woerd H, Wolff E. 1989, A&A 225, 97 Vande Putte D., Smith R.C., Hawkins N.A., Martin J.S. 2001, A&A in press Verbunt F. 1987, A&AS 71, 339 Vogt N., Krzeminski W., Sterken C. 1980, A&A 85, 106 Warner B. 1983, IAU Coll 72, 155 Warner B. 1987, MNRAS 227, 23 Watson M.G., Sherrington M.R., Jameson R.F. 1978, MNRAS 184, 79 Williams R.E. 1980, ApJ 235, 939 Wood M. 1994, IAU Coll. 147, 612
astro-ph/0410491
1
0410
2004-10-20T20:21:56
Bipolar Symbiotic Planetary Nebulae in the Thermal-IR: M2-9, Mz3, and He2-104
[ "astro-ph" ]
We present thermal-IR images of three extreme bipolar objects, M2-9, Mz3, and He2-104. They are bipolar planetary nebulae with bright central stars and are thought to be powered by symbiotic binary systems. The mid-IR images spatially resolve the SEDs of the central engines from the surrounding nebulae. A warm dust component of several hundred degrees can account for the core emission, while a cooler component of about 100 K produces the more extended emission from the bipolar lobes. In every case, the dust mass for the unresolved core region is orders of magnitude less than that in the extended lobes, raising doubts that the hypothetical disks in the core could have been responsible for pinching the waists of the nebulae. We find total masses of roughly 0.5-1 Msun in the nebulae of M2-9 and Mz3, requiring that this material was donated by intermediate-mass progenitor stars. The mass of He2-104's nebula is much lower, and any extended emission is too faint to detect in our images. Extended dust around both M2-9 and Mz3 resembles the distribution of ionized gas. Our images of Mz3 have the highest signal-to-noise in the extended polar lobes, and we show that the fairly uniform color temperature derived from our images can explain the 110 K dust component that dominates the far-IR SED. In the case of Mz3, most of the mass traced by dust is concentrated at high latitudes.
astro-ph
astro-ph
BIPOLAR SYMBIOTIC PLANETARY NEBULAE IN THE THERMAL-IR: M 2-9, Mz 3, and He 2-104 Nathan Smith1,2,3 Center for Astrophysics and Space Astronomy, University of Colorado, 389 UCB, Boulder, CO 80309 and Robert D. Gehrz3 Astronomy Department, University of Minnesota, 116 Church St. SE, Minneapolis, MN 55455 ABSTRACT We present thermal-infrared images of three extreme bipolar objects, M 2-9, Mz 3, and He 2- 104. They are bipolar planetary nebulae with bright central stars and are thought to be powered by symbiotic binary systems. The mid-infrared images spatially resolve the spectral energy distributions of the central engines from the surrounding nebulae. A warm dust component of several hundred degrees can account for the core emission, while a cooler component of ∼100 K produces the more extended emission from the bipolar lobes. In every case, the dust mass for the unresolved core region is orders of magnitude less than that in the extended lobes, raising doubts that the hypothetical disks in the core could have been responsible for pinching the waists of the nebulae. We find total masses of roughly 0.5 -- 1 M⊙ in the nebulae of M 2-9 and Mz 3, requiring that this material was donated by intermediate-mass progenitor stars. The mass of He 2-104's nebula is much lower, and any extended emission is too faint to detect in our images. Extended dust emission is detected around both M 2-9 and Mz 3, in both cases resembling the distribution of ionized gas. Our images of Mz 3 have the highest signal-to-noise in the extended polar lobes, and we show that the fairly uniform color temperature derived from our images can explain the 110 K dust component that dominates the far-infrared spectral energy distribution. In the case of Mz 3, most of the mass traced by dust is concentrated at high latitudes, and we note possible evidence for grain destruction in shocks indicated by an anticorrelation between [Fe ii] and dust emission. Except for these regions with enhanced [Fe ii] emission, the dust continuum closely resembles the distribution of ionized gas. Subject headings: binaries: symbiotic -- circumstellar matter -- planetary nebulae: general -- planetary nebulae: individual (M 2-9, Mz 3, He 2-104) -- stars: mass-loss 1. INTRODUCTION The formation of bipolar structure in planetary nebulae (PNe) is one of the enduring challenges to understanding the late stages of stellar evolution. No simple model has yet emerged that can account for 1Hubble Fellow; [email protected] 2Visiting astronomer at the European Southern Observatory, La Silla, Chile. 3Visiting astronomer at the IRTF, operated by the University of Hawaii under contract with NASA. -- 2 -- how a diffuse slowly-rotating asymptotic giant branch star develops strong bipolarity (not to mention point symmetry, multipolar outflows, jets, etc.) as it becomes a PN, although confining circumstellar disks, disk/jet precession, core rotation, or magnetic fields may play important roles (reviews of shaping mechanisms are given by Balick & Frank 2002; Frank 1999; Mellema & Frank 1995; and contributions in Meixner et al. 2004). One can alleviate the mental strain required to invent self-consistent single-star models by invoking binary systems to provide an axis of symmetry and angular momentum, which are necessary for bipolar phenomena. The "Butterfly" (M 2-9), the "Ant" (Mz 3), and the "Southern Crab" (He 2-104) are among the most dramatic bipolar PNe with tightly-pinched waists seen in Hubble Space Telescope (HST) images (Balick & Frank 2002; Balick 1999; Balick et al. 1997; Corradi et al. 2001). It is interesting that each of these is thought to be powered by a symbiotic binary system (Whitelock 1987; Balick 1989; Corradi & Schwarz 1993; Corradi et al. 2000; Schmeja & Kimeswenger 2001; Smith 2003). These objects have been studied extensively with spectra at visual and near-infrared (IR) wavelengths. They have bright but highly-reddened stellar cores with rich emission-line spectra, accompanied by faint nebular spectra from lower density regions in their polar lobes (Evans 1959; Cohen et al. 1978; Allen & Swings 1972; Swings & Andrillat 1979; Balick 1989; Lutz et al. 1989; Goodrich 1991; Hora & Latter 1994; de Freitas Pacheco & Costa 1996; Phillips & Cuesta 1999; Smith 2003); M 2-9 and Mz 3 in particular are spectroscopic twins at visual and near-IR wavelengths except for the lack of H2 in Mz 3 (Smith 2003). Each object requires an ionization source with a temperature of order 30,000 K, and the nebulae seem to have moderately-enhanced abundances of He and N, suggesting intermediate-mass progenitors. Published distances to each object vary widely, scattered around 1-2 kpc; in this study we quote numerical values like M and L for a nominal distance of 1 kpc. By comparison, little is known about the nature of dust in these sources. The spatial distribution of thermal-IR emission can constrain the dust temperature and mass in the core regions and polar lobes separately, in order to then estimate important quantities like the total mass of the nebulae and the mass of putative circumstellar disks. Low-resolution 8-13 µm spectra of M 2-9 and Mz 3 show similar, nearly flat continuum emission in both sources, except for the presence of [Ne ii] 12.8 µm in Mz 3 (Aitken & Roche 1982), but these observations did not include any spatial information. Although each object was observed with the Infrared Astronomical Satellite (IRAS), the low spatial resolution of those data cannot separate the relative contributions of the core and polar lobes. Ground-based images in the thermal-IR with arcsecond- scale spatial resolution have only been published for Mz 3 (Quinn et al. 1996), but that was one image in a single broad filter, making it difficult to constrain the dust temperature or mass. In an effort to clarify the spatial distribution of mid-IR properties of M 2-9, Mz 3, and He 2-104, we obtained 8-25 µm images of these sources with MIRLIN on the IRTF and TIMMI2 on the ESO 3.6m telescope. 2. OBSERVATIONS 2.1. MIRLIN Images of M 2-9 We obtained thermal-IR images of M 2-9 on the nights of 2001 July 22, 23, and 24 using the 5 -- 26 µm camera MIRLIN mounted on the 3m NASA Infrared Telescope Facility (IRTF). In this configuration, MIRLIN's 128×128 Si:As BIB array has a pixel scale of 0.′′465 and a field of view of 1′. Chop-nod images were obtained in three broadband filters, N1, Qs, and Q5 centered at 8.8, 17.9, and 24.5 µm, respectively. We also used the 2% resolution circular variable filter (CVF) centered at 10.5 and 12.8 µm to image the emission lines of [S iv] and [Ne ii], respectively (see Table 1). The weather conditions were not optimal (intermittent thin clouds) on the first night when the 12.8 µm image was obtained, and the photometric uncertainty of that -- 3 -- image is ±30%, but the weather was better on the next two nights when the other images were obtained, and for those the photometric uncertainty is 10-20% (larger uncertainties in the Q-band images). The chopper throw was 30′′ east/west with a similar north/south telescope nod that we varied slightly from one set of images to the next. After sky subtraction, the many individual frames were resampled to a smaller pixel scale and then shifted and added, using the bright central star for spatial registration. The observations were flux calibrated using observations of α Sco, adopting the zero magnitude fluxes in the MIRLIN handbook. Figure 1 shows the final flux-calibrated images of M 2-9 in the N1, Qs, and Q5 filters, where the lowest contour is drawn at approximately 3σ above the background. The [S iv] 10.5 µm and [Ne ii] 12.8 µm images are not displayed in Figure 1 because only the central point source and no extended structure was detected. Even in the continuum images, the extended structure in the nebula is very faint and only marginally detected, so we have not made ratio maps displaying the color temperature and emitting optical depth. The faint mid-IR extended emission from warm dust is clearly elongated in the north-south direction, similar to the structure seen in HST and near-IR images (Balick & Frank 2002; Hora & Latter 1994). This extended structure is discussed in more detail below in §4. Table 1 gives sky-subtracted flux densities for the central star in each filter, measured in a 4′′ diameter circular software aperture. 2.2. TIMMI2 Images of Mz 3 and He 2-104 On 2003 May 15 we used the Thermal Infrared Multi-Mode Instrument (TIMMI2) on the ESO 3.6m telescope at La Silla, Chile to obtain mid-IR images of Mz 3 and He 2-104. TIMMI2 has a 240×340 pixel Si:As BIB detector with a pixel scale of 0.′′2 and a field of view of 48′′×64′′. Chop-nod images were obtained in three broadband filters centered at 8.9, 11.9, and 17.0 µm, and with the [Ne ii] 12.8 µm narrowband filter (see Table 2). Different chop-nod patterns were used for each source, due to the different sizes of the nebulae in optical images. The polar lobes of Mz 3 have a spatial extent of roughly 20′′×30′′ in emission-line images (Lopez & Meaburn 1983; Redman et al. 2000; Smith 2003) allowing us to place the positive and negative images of the nebula side-by-side on the array. Therefore, to observe Mz 3 we used a chopper throw of 40′′ north-south (chopping off the array) and east-west nods of 30 -- 35′′. Since the inner bipolar lobes or "rings" of He 2-104 are only ∼10′′ across in optical images (Corradi et al. 2001), we used a chop-nod pattern appropriate for a point source, where we chopped 18′′ north-south and nodded the telescope 20′′ east-west. This allowed the target to be placed on the array four times for each set of chop-nod observations, increasing the effective on-source exposure time. For flux calibration, we obtained similar observations of γ Cru on the same night and at similar airmass, using fluxes for each filter listed in the TIMMI2 user manual. The weather was mostly photometric, and the photometric accuracy in our images is 5 -- 10%, dominated by uncertainty in the standard star calibration. Figures 2 and 3 show the resulting TIMMI2 images of Mz 3 and He 2-104, respectively, where the lowest contour is 3σ above the background. Table 2 gives sky-subtracted flux densities for the central star in each filter, measured in a 4′′ diameter circular software aperture for both Mz 3 and He 2-104. Figure 2 reveals obvious extended thermal-IR emission from warm dust in the bipolar lobes around Mz 3 in all three continuum filters, as well as strong [Ne ii] emission. The observed morphology is consistent with the pair of polar bubbles along the nearly north-south axis seen in emission-line images at shorter wavelengths (Smith 2003; Santander-Garcia et al. 2004; Guerrero et al. 2004). This extended structure is discussed in §5. Our images of He 2-104 had even better sensitivity than those of Mz 3, but were not sensitive enough to detect significant extended structure from the bipolar lobes or rings of He 2-104. The 11.9 and -- 4 -- 12.8 µm images in Figure 3 do show a very faint hint of extended structure, with the lowest contour slightly elongated along a northeast/southwest direction out to ∼3′′ from the star. Elongation along this direction (which is different from both the chop and nod directions) would be consistent with the elongated central emission structure seen in HST images (Corradi et al. 2001), marking the limb-brightened section near the equator where the two polar lobes appear to meet. Our images demonstrate that the extended mid-IR structure of He 2-104 is much fainter compared to the central source than in either M 2-9 or Mz 3. 3. PHOTOMETRY OF THE CENTRAL ENGINES All three of our targets show a bright, unresolved central source at mid-IR wavelengths, related to their status as potential symbiotic binaries. This is very different than the case of η Carinae, to which these three planetary nebulae are often compared, where high-resolution IR images reveal a complex group of dust clumps arranged in a disrupted torus at the point where the two polar lobes meet at the equator (Smith et al. 2002). Our images of both M 2-9 and Mz 3 reveal extended emission from warm dust in the bipolar lobes of each object, while we detect no extended mid-IR emission around He 2-104 at comparable sensitivity. However, spatially-resolved mid-IR photometry of these central engines reveals important clues to the nature of all three objects. The extended nebulae of M 2-9 and Mz 3 are discussed separately in §4 and §5. 3.1. Dust Temperature, Luminosity, and Mass Figure 4 displays the spectral energy distributions (SEDs) of the bright central cores measured from our images using a 4′′ diameter aperture (from Tables 1 and 2), compared to the integrated emission in a larger aperture over a wider range of IR wavelengths measured by 2MASS, MSX, and IRAS.1 For each source, a solid curve shows an approximate gray-body fit to the SED of the unresolved central engine, which excludes extended emission from dust in the nebulae. In all three cases, the integrated emission at longer IR wavelengths clearly exceeds this solid curve, and the dotted curve shows the contribution from additional cool dust that must be emitted from a larger area on the sky than the aperture we used to measure each of the central stars (the dashed curve shows the sum of both, which fits the total 8 -- 100 µm SED for each object). We take these two dust temperature components to represent warm dust in the unresolved central core and cool dust in the more extended nebulae. Temperatures, luminosities, and dust masses for these warm and cool dust components are listed in Table 3, calculated for a nominal distance of 1 kpc. The dust mass for each temperature component in Table 3 was derived from the total integrated lumi- nosity of the individual curves drawn in Figure 4, with necessary assumptions about the grain properties. The relation for the mass of dust required to account for the observed IR luminosity can be expressed as Md = h aρ 3σQeT 4 d iLd (1) where a, ρ, and Qe are the effective grain radius, mass density, and mean thermal emissivity of the optically important grains, and σ is the Stefan-Boltzmann constant. Since no silicate emission features are seen in any of the three objects we observed, we assume that carbon grains dominate the IR emission, and we adopt 1These data were collected from the Infrared Science Archive at http://irsa.ipac.caltech.edu/. -- 5 -- ρ ≃2.25 g cm−3. We do not know the grain size, but since the recurrent outbursts that may form symbiotic planetary nebulae are likely caused by nova-like outbursts, we assume a .0.2 µm (i.e., the maximum grain size observed in novae; Gehrz 1999). With grains smaller than 0.2 µm, the grain emissivities given by Gilman (1974) can be approximated with which permits us to write an expression for the dust mass that is independent of the grain emissivity and radius. We have Qe = 1 100 aT 2 d (2) Md = h 100ρ 3σT 6 d iLd (3) which is used to calculate the dust mass for each component listed in Table 3 and discussed below for each object. Given the inherent uncertainties in the flux calibration and uniqueness of the fits to the IR data, the dust masses we calculate are probably uncertain at the ±30% level. Table 3 also lists likely total masses (gas + dust) assuming a gas-to-dust mass ratio of 230, which is appropriate if the grains fully deplete the carbon in solar-composition ejecta (Grevesse & Anders 1989). We presume that at shorter near-IR wavelengths, the continuum SED of each object is dominated by the unresolved central source (e.g., Hora & Latter 1994; Smith 2003). At those wavelengths, the SED is an uncertain combination of reddened photospheric emission from the central stars, hot dust in the core, and scattered light from the lobes, so we do not show a fit to the 2MASS data in Figure 4. Even if extended hot dust in the polar lobes made a significant contribution at these wavelengths, the total mass would be negligible compared to the mass of cooler dust (e.g., Smith et al. 1998; and Table 3). In any case, the very simple fits in Figure 4 are sufficient to characterize the dust emission that dominates in the mid-IR. 3.2. Mz 3 Of the three targets we observed, Mz 3 has the smallest fraction of the total emission contributed by its unresolved core. At a wavelength of ∼12 µm, for example, the central source contributes less than half the total flux measured by MSX and IRAS. The relative contribution of the core weakens toward longer wavelengths, supplying ∼10% at 20 µm, and only about 1% at 60 µm. The integrated IR luminosity is only about 25% of the total presumed bolometric luminosity of L=104 L⊙ (Smith 2003), indicating that the dust is optically thin to the escaping UV and visual-wavelength stellar radiation over most of the solid angle seen by the central engine. The mass of dust in the central core is difficult to guage accurately from the present data, because the emission may be optically thick. However, the dust mass needed to emit the optically thin 320 K gray-body shown in Figure 4 gives a useful lower estimate of the mass required in the outer part of the disk. This core mass for the warm dust component is negligible compared to the much larger total mass of ∼0.6 M⊙ (gas + dust) for the cooler material farther from the star in the bipolar lobes. This relatively large mass for the circumstellar ejecta, which approaches 1 M⊙, reinforces the conjecture from chemical abundances that the bipolar lobes were ejected from an intermediate-mass progenitor star (Smith 2003). Our imaging photometry reveals a significant 12.8 µm [Ne ii] emission line from the central source, in -- 6 -- agreement with spectra taken by Aitken & Roche (1982). We measure a continuum-subtracted flux of ∼9 Jy or 3.6×10−11 ergs s−1 cm−2 (integrated over the filter bandpass), and a corresponding equivalent width of ∼530 A. The observed mid-IR SED of the unresolved central source is somewhat flatter than a single gray- body, implying the existence of a range of dust temperatures. Thus, the single temperature of 320 K is only representative for the purpose of calculating the minimum luminosity and mass of the dust (Table 3). Indeed, an additional component of hot dust at ∼900 K is required to fit the near-IR continuum spectrum (Smith 2003; Cohen et al. 1978). This wide range of temperatures from the unresolved central source probably indicates the existence of a circumstellar disk with dust located at a range of different radii. The observed properties of the core SED allow us to place some constraints on the hypothetical disk. The 320 K component represents the cooler, outer parts of the disk, while the 900 K dust (Smith 2003) traces the inner region of the disk, near the dust sublimation radius. For L=104 L⊙, these correspond to 10 and 80 AU, repsectively, or 10 to 80 mas for D=1 kpc. In any case the disk is much smaller than the spatial resolution of our images and near the limit attainable with HST, but within the current reach of IR interferometry. The disk is evidently very thin; the disk's IR luminosity is less than about 5% of the total available bolometric luminosity (Smith 2003), indicating a half opening angle of .12◦. 3.3. M 2-9 The SED of M 2-9 shown in Figure 4a is similar to that of Mz 3, except that the relative contribution from extended cool dust is weaker. The 8 -- 20 µm SED of the central source is almost identical to Mz 3, both having an N-band flux of ∼30 Jy and an intrinsic IR luminosity just below 500 L⊙; however, M 2-9 lacks significant 12.8 µm [Ne ii] emission (Aitken & Roche 1982). The 260 K fit to the SED of the central source is very approximate, since the observed SED is actually flatter than any single temperature, perhaps indicating dust at various radii in a circumstellar disk. As with Mz 3, a hotter dust component is needed to account for the near-IR continuum (Hora & Latter 1994). The central source also requires dust at a somewhat cooler average temperature compared to Mz 3, implying that the central source is intrinsically less luminous or that the dust is distributed toward larger radii in the disk. The cooler dust component that dominates at far-IR wavelengths, emitted by the extended dust in the polar lobes of M 2-9, supplies a smaller fraction of the total IR luminosity than its counterpart in Mz 3. Also like Mz 3, the warm dust in the core of M 2-9 makes a negligible contribution to the total mass of the circumstellar nebula. And finally, like Mz 3, the relatively large total mass we derive for the cooler component in the extended nebula (about 0.8 M⊙) requires that the progenitor that ejected the mass was probably an intermediate-mass star of at least a few solar masses. Altogether, the similarities between the IR properties of M 2-9 and Mz 3 are remarkable. Since H2 formation is linked to dust grains in most circumstances, these similarities underscore the mystery of why the near-IR H2 lines are so prominent in the polar lobes of M 2-9 (Hora & Latter 1994), while being totally absent in Mz 3 (Smith 2003). -- 7 -- 3.4. He 2-104 All our targets are dominated by a bright unresolved core, but He 2-104 is the only one in our sample showing no clear evidence for extended emission at 8 -- 12 µm. The bright unresolved central engine of He 2- 104 dominates the SED at all IR wavelengths we have observed, and the core flux we measure matches the integrated 12 µm IRAS flux. He 2-104 is also significantly fainter than the other two sources; the warm dust component has less than 1/3 of the N-band flux or bolometric IR luminosity of M 2-9 or Mz 3. The SED of He 2-104 shows no evidence for any 12.8 µm [Ne ii] emission. In addition to the 320 K dust component that dominates at mid-IR wavelengths, He 2-104 also requires a component of hotter dust to account for the observed near-IR photometry, as is the case for both M 2-9 and Mz 3. Like the other two, the addition of a cooler dust component is needed at far-IR wavelengths to explain the integrated IRAS fluxes. However, in He 2-104 the cool 95 K component is far less luminous than the other two: it has only 30 L⊙, and it is the only source in which the luminosity of the cool dust component at far-IR wavelengths is much lower than the warmer dust component that dominates in the mid-IR. From equation 3, the mass of dust required to emit the observed far-IR flux from the 95 K component in He 2-104 is only 10−4 M⊙, or a total gas + dust mass of ∼0.02 M⊙. This is a much smaller mass than either M 2-9 or Mz 3, which explains why we did not detect any extended structure in our images, but does not let us place strong constraints on the likely mass of the progenitor star. 3.5. Where's the Donut? Here we briefly note that in all three objects, the mass of ejecta in the unresolved central core/disk is more than 100 times lower than the mass of ejecta in the more extended polar lobes. This fact is relevant for the quest to understand the origin of bipolar structure (see Balick & Frank 2002), since with disk masses that are orders of magnitude less than the polar lobes, these disks cannot provide the resistance needed to absorb momentum and thereby constrict the equatorial expansion of an otherwise spherical shell. The same problem exists for η Carinae as well, where the hot equatorial dust in the core region provides negligible mass compared to the polar lobes (Smith et al. 2003; Frank et al. 1998). Some other mechanism that favors polar ejection of material may be required (see e.g., Balick & Frank 2002; Matt & Balick 2004; contributions in Meixner et al. 2004). Observations of stars still on the asymptotic giant branch suggest that asymmetries in proto-planetary nebulae start early, either through rotation or binary interactions (Gehrz 2004). 4. EXTENDED STRUCTURE OF M 2-9 Figure 1 reveals clear extended structure in the bipolar lobes of M 2-9, extending out to 15 -- 20′′ from the star toward both the north and south, consistent with the extent of the bipolar lobes seen in HST and IR images (Balick & Frank 2002; Hora & Latter 1994). Of the three wavelengths shown in Figure 1, the highest quality image is Figure 1b at 18 µm. In this image, the right (west) side of the nebula is clearly brighter than the left. This is the opposite of the case in emission-line tracers of ionized gas (Balick & Frank 2002; Hora & Latter 1994), which are much brighter on the east side of the nebula; note, however, that this pattern changes with time (Doyle et al. 2000; Allen & Swings 1972). This difference between emission lines and dust emission suggests a spatial anticorrelation between ionized gas and warm dust in the variable ionization structure of M 2-9's bipolar lobes. Perhaps temporary exposure to the hard UV radiation field of the central source is sufficient to destroy significant quantities of dust. The destruction of dust grains by -- 8 -- UV radiation is supported by observations of novae (Gehrz et al. 1980a, 1980b). In Figure 1, the polar lobes of M 2-9 appear somewhat thicker at 24.5 µm than they do at 8 -- 18 µm. A larger size at longer IR wavelengths might imply the existence of an outer shell of cooler dust that emits less efficiently at shorter IR wavelengths. Deeper images with higher spatial resolution are obviously needed to confirm this conjecture. Although poorly justified by the quality of the images in Figure 1, the possible existence of a double shell structure -- a cool outer dust shell and a warmer inner dust shell -- is motivated by other observational clues. M 2-9 is frequently compared to the Homunculus nebula around η Car, which shows precicely this type of double-shell structure in thermal-IR images (Smith et al. 2003). A double-shell morphology is reinforced by the near-IR emission line structure of η Car, with H2 delineating a thin outer shell, and a smaller inner shell seen clearly in [Fe ii] lines (Smith 2002). This is identical to the near-IR excitation structure observed in these same emission lines in images of M 2-9 (Hora & Latter 1994). In addition to the similarities in the dust temperature and nebular excitation structure, M 2-9 and η Car share other observational characteristics. Several observers have noted similarities in the very rich emission-line spectra of the central sources, including strong [Fe ii] emission (e.g., Balick 1989; Swings & Andrillat 1979; Allen & Swings 1972). More recently, it has been recognized that they also share a similar and very rare type of temporal variability in their nebulae. M 2-9 shows bizarre changes in its apparent brightness distribution, which are usually attributed to an azimuthally-evolving UV radiation field caused by moving shadows from the cooler component in the central binary system (Doyle et al. 2000; Allen & Swings 1972). Multiepoch UV images of η Car have recently revealed a similar type of variability in its "Purple Haze", with an anlogous root cause (Smith et al. 2004). Finally, Figure 1 (especially Fig. 1b) shows emission from the knot S3, and possibly its northern coun- terpart N3 as well, both at roughly 15′′ from the star. This is significant, because it may indicate that dust survives in these condensations, which are thought to be shock excited. They are presumably formed by shocks in a fast polar wind or jet, and are sometimes called FLIERS (fast low-ionization emission regions; Balick & Frank 2002) or ansae. Their bright [Fe ii] emission is thought to be due to the liberation of Fe into the gas phase following the destruction of dust. Thus, their detection in dust emission at thermal-IR wavelengths is significant. However, [Fe ii] might contaminate the 17.9 µm filter. 5. EXTENDED STRUCTURE OF Mz 3 5.1. Morphology in Images Of our three targets, Mz 3 shows the brightest and most dramatic extended structure. Although Mz 3 has a complex ejecta pattern with multiple polar lobes seen in images (Smith 2003; Santander-Garcia et al. 2004; Guerrero et al. 2004), here we are only concerned with the innermost bipolar lobes (i.e. the head and abdomen of the ant). Viewing the images on a computer display shows considerably more detailed structure than is conveyed in Figure 2, so we applied 20 iterations of the lucy deconvolution algorithm in IRAF to our 11.9 µm continuum and 12.8 µm [Ne ii] images, which have the best signal-to-noise in our dataset. The results are shown in Figures 5a and 5b, respectively.2 At any position in the polar lobes, the [Ne ii] intensity is significantly higher than the continuum emission -- from Figure 5 we have typically 2The emission spot immediately west of the central star in Figure 5a is an artifact of the sky subtraction offset settling time, as opposed to real structure. Our standard star used as the PSF in the deconvolution had a slightly different chop-nod pattern. -- 9 -- Integrating over the 2200 A filter bandpass, this indicates a typical equivalent Iλ([Ne ii]) ≃ 3 × Iλ(11.9). width in the lobes of ∼4400±400 A, which is much higher than in the central star because of the cooler and optically thinner dust continuum (see below). Most of the extended 11.9 µm emission -- and therefore most of the dust mass -- appears to be concen- trated in caps at the top and bottom of the polar lobes, rather than in their side walls. If the dust mass traces the gas mass, then this is a crucial fact as it bears on the latitudinal distribution of mass in the initial ejection. There is good reason to think that the dust mass does generally trace the total gas mass in the lobes, since the morphology of the 11.9 µm image is almost identical to the [Ne ii] images, as well as nar- rowband emission-line images at shorter wavelengths (Smith 2003; Santander-Garcia et al. 2004; Guerrero et al. 2004). On the other hand, the regions of the polar lobes where the dust appears to be deficient -- in the side walls at low latitudes and in the polar "blisters" beyond the lobes -- are precisely the locations of low excitation where infrared [Fe ii] emission is mysteriously enhanced (Smith 2003). In other words, there is an anti-correlation between dust and [Fe ii] emission, which might suggest that some of the excess [Fe ii] emission has resulted as iron atoms were liberated from grains into the gas phase by shocks. We also point out that the polar "blisters" are seen in the [Ne ii] image, especially to the north. Interestingly, the dust morphology resembles the high-excitation emission tracers like He i and [O iii] somewhat better than it does hydrogen lines. For example, the missing side walls of the polar lobes are a characteristic of both Figure 5a and the He i λ10830 emission (Smith 2003), while the limb-brightened side walls are seen clearly in hydrogen and [Fe ii] lines. This is somewhat surprising, since dust and high- excitation emission from ionzed gas are often segregated (the nebula of RY Scuti is a clear example; Gehrz et al. 2001; Smith et al. 2001). The similarity of the warm dust emission and ionized gas may have implications for the heating mechanism (i.e. either they are heated radiatively by the same latitudinally-dependent UV field, or perhaps the dust is heated by trapped Lyα photons in the nebula, etc.). Finally, several authors have noted excess extinction at low latitudes (Smith 2003; Guerrero et al. 2004) or evidence for dust in the side walls of the polar lobes from polarization data (Scarrott & Scarrott 1995), while the majority of the dust in our images is located in the polar caps as noted above. This indicates that the extra dust that may be causing this extinction and polarization in scattered light may be cold, with insufficient heating to produce detectable mid-IR emission. 5.2. Dust Temperature and Optical Depth Figures 5c and 5d show the spatial distribution of dust color temperature and emitting optical depth in the polar lobes of Mz 3. The color temperature for each pixel in Figure 5c is given by Tc = 14404 [(1/λ2) − (1/λ1)] ln{[Fν(λ1)/Fν(λ2)] (λ1/λ2)β+3} K (4) where Fν (λ1)/Fν(λ2) is the flux ratio between the continuum at λ1 (11.9 µm) and λ2 (17 µm), and −β is the dust emissivity exponent. We have assumed that at these wavelengths the grains have an emissivity proportional to λ−1 (i.e., β=1).3 To calculate this color temperature image, we clipped the input 11.9 µm image at a level of 0.18 Jy arcsec−2, with lower values set to zero. The resulting color temperature map was 3Note that if this assumed emissivity is wrong, then smaller values of β would result in higher derived color temperatures. -- 10 -- then used to calculate the distribution of the emitting optical depth τ of warm grains in the lobes of Mz 3 (note that τ is not the absorption optical depth). The emitting optical depth at some wavelength is given by τ = − lnh1 − Iν Bν(Tc)i (5) where Iν is the specific intensity in an input image, and Bν(Tc) is the Planck function corresponding to a given color temperature in each pixel of the Tc image. The apparent dust temperature distribution in Mz 3's polar lobes in Figure 5c is fairly uniform, dom- inated by temperatures of 100 -- 130 K, while peak temperatures in the unresolved core region exceed 300 K. Thus, the spatially-resolved temperature structure confirms our interpretation of the SED in Figure 4b. Consequently, the "cool" mass estimates we list in Table 3 are reliable values for Mz 3's nebula, without contamination by dust in the central disk. The τ map for Mz 3 in Figure 5d shows values that typically range from 0.01 to 0.05, so it is unlikely that a significant amount of mass is hidden in optically thick clumps. Only in the brightest part of the polar lobes, roughly 8′′ northwest of the star, do we see optical depths that exceed 0.1. The observed grain color temperature and the apparent separation from the central engine provide a consistency check for evaluating our assumptions about distance, luminosity, and grain properties. Most of the dust in the polar lobes appears to reside at a separation of ∼10′′ or R=104 AU if the distance to Mz 3 is roughly 1 kpc. The most likely bolometric luminosity of the central engine is of order 104 L⊙ (Smith 2003). Then, the grain temperature should correspond to Tc = 28h Qabs Qe L 104L⊙ 104AU(cid:17)−2i (cid:16) R 1 4 K (6) where Qabs/Qe is the ratio of absorption to emission efficiency for the grains. For blackbodies (i.e., Qabs/Qe=1) we would expect a dust temperature around 30 K, much lower than observed. The higher observed temperature would require Qabs/Qe ≃200, which would be consistent with small graphite grains with radii of a ≃0.1 µm (Gilman 1974). This is consistent with the roughly λ−1 wavelength-dependence that we assumed initially to calculate the dust temperature. Alternatively, if the grains are larger than 0.1 µm (i.e. a somewhat smaller value of Qabs/Qe), then the bolometric luminosity could be somewhat larger or the distance smaller than we assumed. 6. CONCLUSIONS We presented thermal-IR images of the bipolar planetary nebulae M 2-9, Mz 3, and He 2-104, obtained at the IRTF and at ESO. The distribution of dust in these systems is of interest as each object is strongly bipolar with a bright central star, and each is thought to be powered by a symbiotic binary system. Our mid-IR images allowed us to construct SEDs of the bright, unresolved central engines separate from the surrounding nebulae. In each case, we find that a warm dust component of several hundred degrees can account for the core emission (260, 320, and 320 K in M 2-9, Mz 3, and He 2-104, respectively), while a cooler component produces the more extended emission from the bipolar lobes (85, 110, and 95 K in M 2-9, Mz 3, and He 2-104, respectively). Hot dust at temperatures up to ∼900 K is needed to fit the near-IR SEDs of each source, but these additional components contribute negligible mass. -- 11 -- In every case, the dust mass for the unresolved core region is orders of magnitude less than the mass of the extended lobes, raising doubts that the hypothetical disks in the core could have been responsible for pinching the waists of the bipolar nebulae. Assuming a gas:dust mass ratio of 230 appropriate for graphite grains and solar composition, we find masses of roughly 0.8 and 0.6 M⊙ in the nebulae of M 2-9 and Mz 3, respectively. These relatively high masses require that this material was donated by intermediate-mass progenitor stars. The mass of He 2-104's nebula is much lower, in accordance with the fact that we detect no extended dust emission from its much fainter nebula. We detected extended dust emission around both M 2-9 and Mz 3, and in both cases it resembled the distribution of ionized gas seen in narrowband emission-line images. Our images of Mz 3 have the highest signal-to-noise in the extended polar lobes, and we show that the fairly uniform color temperature derived from these images can explain the 110 K dust component that dominates the far-IR spectral energy distribution. Thus, the mass derived from the cool 110 K component traces material in the polar lobes. In the case of Mz 3, the dust is concentrated at high latitudes. We also note possible evidence for grain destruction in shocks, indicated by an anticorrelation between dust seen in our images and [Fe ii] emission in previously-published near-IR images. While dust appears to be absent or weak in these regions (i.e. the polar "blisters" of Mz 3), 12.8 µm [Ne ii] emission is detected. Except for these regions with enhanced [Fe ii] emission, the thermal-IR continuum from dust closely resembles the distribution of ionzed gas. Given the dust temperatures of the as-yet unresolved central sources, where dust presumably resides in a circumstellar disk, each of these bright objects may prove to be worthwhile targets for mid-IR interferometric observations to resolve the disks. N.S. was supported by NASA through grant HF-01166.01A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. R.D.G. was supported by NASA, the NSF, and the United States Air Force. Aitken, D.K., & Roche, P.F. 1982, MNRAS, 200, 217 Allen, D.A., & Swings, J.P. 1972, ApJ, 174, 583 Balick, B. 1989, AJ, 97, 476 REFERENCES Balick, B. 1999, in ASP Conf. Ser. 188, Optical and Infrared Spectroscopy of Circumstellar Matter, eds. E.W. Guenther, B. Stecklum, & S. Klose (San Francisco: ASP), 241 Balick, B., & Frank, A. 2002, ARAA, 40, 439 Balick, B., Icke, V., & Mellema, G. 1997, HST Press Release, STScI-PR97-35 Cohen, M., Fitzgerald, M.P., Kunkel, W., Lasker, B.M., & Osmer, P.S. 1978, ApJ, 221, 151 Corradi, R.L.M., & Schwarz, H.E. 1993, A&A, 268, 714 Corradi, R.L.M., et al. 2000 Corradi, R.L.M., et al. 2001 Doyle, S., Balick, B., Corradi, R.L.M., & Schwarz, H.E. 2000, AJ, 119, 1339 de Freitas Pacheco, J.A., & Costa, R.D.D. 1996, A&A, 309, 629 Evans, D.S. 1959, MNRAS, 119, 150 Frank, A. 1999, New Astron. Rev., 43, 31 Frank, A., Ryu, D., & Davidson, K. 1998, ApJ, 500, 291 Gehrz, R.D. 1999, in Processes in Astrophysical Fluids, eds. O. Regev & D. Prialnik, Physics Reports, 311, 405 Gehrz, R.D., Smith, N., Jones, B., Puetter, R., & Yahil, A. 2001, ApJ, 559, 395 -- 12 -- Gehrz, R.D., Grasdalen, G.L., Hackwell, J.A., & Ney, E.P. 1980a, ApJ, 237, 855 Gehrz, R.D., Hackwell, J.A., Grasdalen, G.L., Ney, E.P., Neugebauer, G., & Sellgren, K. 1980b, ApJ, 239, 570 Gehrz, R.D. 2004, in ASP Conf. Ser. 313, Asymmetrical Planetary Nebulae III, Meixner, M., Kastner, J.H., Balick, B., & Soker, N., eds. (San Francisco: ASP), 327 Gilman, R.C. 1974, ApJS, 28, 397 Goodrich, R.W. 1991, ApJ, 366, 163 Grevesse, N., & Anders, E. 1989, in Cosmic Abundance of Matter, ed. C.J. Waddington (AIP: New York), 1 Guerrero, M.A., Chu, Y.H., & Miranda, L.F. 2004, AJ, 128, 1694 Hora, J.L., & Latter, W.B. 1994, ApJ, 437, 281 Lopez, J.A., & Meaburn, J. 1983, MNRAS, 204, 203 Lutz, J.H., Kaler, J.B., Shaw, R.A., Schwarz, H.E., & Aspin, C. 1989, PASP, 101, 966 Matt, S., & Balick, B. 2004, ApJ, in press Meixner, M., Kastner, J.H., Balick, B., & Soker, N. (eds.) 2004, ASP Conf. Ser. 313, Asymmetrical Planetary Nebulae III (San Francisco: ASP) Mellema, G., & Frank, A. 1995, MNRAS, 273, 401 Phillips, J.P., & Cuesta, L. 1999, AJ, 118, 2919 Quinn, D.E., Moore, T.J.T., Smith, R.G., Smith, C.H., & Fujiyoshi, T. 1996, MNRAS, 283, 1379 Redman, M.P., O'Connor, J.A., Holloway, A.J., Bruce, M., & Meaburn, J. 2000, MNRAS, 312, L23 Santander-Garcia, M., Corradi, R.L.M., Balick, B., & Mampaso, A. 2004, A&A, 426, 185 Scarrott, S.M., & Scarrott, R.M.J. 1995, MNRAS, 277, 277 Schmeja, S., & Kimeswenger, S. 2001, A&A, 377, L18 Smith, N. 2002, MNRAS, 337, 1252 Smith, N. 2003, MNRAS, 342, 383 Smith, N., Gehrz, R.D., & Krautter, J. 1998, AJ, 116, 1332 Smith, N., Gehrz, R.D., & Goss, W.M. 2001, AJ, 122, 2700 Smith, N., Gehrz, R.D., Hinz, P.M., Hoffmann, W.F., Mamajek, E.E., Meyer, M.R., & Hora, J.L. 2002, ApJ, 567, L77 Smith, N., Gehrz, R.D., Hinz, P.M., Hoffmann, W.F., Hora, J.L., Mamajek, E.E., & Meyer, M.R. 2003, AJ, 125, 1458 Smith, N., Morse, J.A., Collins, N.R., & Gull, T.R. 2004, ApJ, 610, L105 Swings, J.P., & Andrillat, Y. 1979, A&A, 74, 85 Whitelock, P.A. 1987, PASP, 99, 573 This preprint was prepared with the AAS LATEX macros v5.2. Table 1. IRTF/MIRLIN Observations of M 2-9 Filter N1 [S iv] [Ne ii] Qs Q5 λ (µm) ∆λ (µm) Exp. Time (sec.) 8.81 10.5 12.8 17.9 24.5 0.87 2% 2% 2.0 0.76 2700 5300 3900 6200 2700 a Fν (Jy) 30.3 32.8 41.1 41.5 38.1 aCentral star in a 4′′ diameter aperture. -- 13 -- Fig. 1. -- IRTF/MIRLIN images of M 2-9 in continuum filters centered at 8.8 µm (a), 17.9 µm (b), and 24.5 µm (c). Images of the [S iv] and [Ne ii] emission lines obtained with the CVF are not shown because no extended structure was detected. The lowest contour level is drawn at the 3σ level above the background, and contour levels in Jy arcsec−2 are listed in the lower right part of each frame. -- 14 -- Fig. 2. -- TIMMI2 images of Mz 3 in the 8.9 µm continuum (a), in the 11.9 µm continuum (b), in the 12.8 µm [Ne ii] filter (c), and in the 17 µm continuum (d). In each panel, the lowest contour level is 3σ above the background, and contour levels in Jy arcsec−2 are listed in the lower right. -- 15 -- Table 2. TIMMI2 Observations of Mz 3 and He 2-104 Filter N8.9 N11.9 [Ne ii] Q1 λ (µm) ∆λ (µm) 8.7 11.6 12.8 17.0 0.78 1.20 0.22 0.80 Exp. Time (Mz3) Fν (Mz3) Exp. Time (He2) Fν (He2) (sec.) 720 1200 1080 1200 (Jy) 34.3 36.9 45.8 45.7 (sec.) 2880 2880 4320 · · · (Jy) 8.1 8.8 8.7 · · · Note. -- Fν for each source corresponds to the bright central star measured through a 4′′ diameter circular aperture. Table 3. IR Luminosity and Dust Mass Parameter units M 2-9 Mz 3 He 2-104 a 260 470 320 462 320 106 a a 4.0×10−6 9.2×10−4 1.1×10−6 2.5×10−4 2.6×10−7 6.0×10−5 Tc warm dust K L warm dust L⊙ M warm dust M⊙ M⊙ M warm gas K Tc cool dust L cool dust L⊙ M⊙ M cool dust M cool gas M⊙ a 85 483 110 1770 95 29 a a 3.4×10−3 2.6×10−3 1.0×10−4 0.78 0.60 0.02 aAt a distance of 1 kpc. Fig. 3. -- TIMMI2 images of He 2-104 in continuum emission at 8.9 µm (a), 11.9 µm (b), and 17 µm (c). In each panel, the lowest contour level is 3σ above the background, and contour levels in Jy arcsec−2 are listed in the lower right. -- 16 -- Fig. 4. -- Spectral energy distributions of M 2-9, Mz 3, and He 2-104 in Panels a, b, and c, respectively. Imaging photometry of the central stars from Tables 1 and 2 is shown with solid circles, while 2MASS, MSX, IRAS photometry are plotted with unfilled squares, crosses, and unfilled triangles, respectively. The solid line is an approximate fit to the mid-IR photometry of the central star using blackbodies with λ−1 emissivity, the dotted line is a similar fit to the integrated far-IR photometry, and the dashed line shows the sum of both. Warmer components are required to fit the 2MASS photometry, but these are not shown as they contribute negligible mass. The integrated fluxes including all extended structure detected in our images of M 2-9 and Mz 3 are consistent with the total MSX and IRAS fluxes at similar wavelengths, but are not shown in this plot because of the large uncertainties. -- 17 -- Fig. 5. -- Panels (a) and (b) show lucy-deconvolved images of Mz 3 in the 11.9 µm continuum and in the 12.8 µm [Ne ii] line. Contour levels in Jy arcsec−2 are listed in the lower right corner of each panel. (c) 11.9 -- 17 µm color temperature in the lobes of Mz 3. The input 11.9 µm image was clipped at 0.18 Jy arcsec−2 and all values below that set to zero, and the 11.9 µm image was smoothed to match the spatial resolution of the 17 µm image. Contours in Kelvin are listed in the lower right. (d) Optical depth image at 11.9 µm using the temperature image in Panel (c), with optical depth contours listed in the lower right.
astro-ph/9807092
1
9807
1998-07-09T13:28:35
Large-Scale Mass Power Spectrum from Peculiar Velocities
[ "astro-ph" ]
This is a brief progress report on a long-term collaborative project to measure the power spectrum (PS) of mass density fluctuations from the Mark III and the SFI catalogs of peculiar velocities. The PS is estimated by applying maximum likelihood analysis, using generalized CDM models with and without COBE normalization. The application to both catalogs yields fairly similar results for the PS. The robust result is a relatively high PS, with P(k)\Omega^{1.2}=(4.5+/-2.0)X10^3 (Mpc/h)^3 at k=0.1 h/Mpc. An extrapolation to smaller scales using the different CDM models gives \sigma_8\Omega^{0.6}=0.85+/-0.2. The general constraint on the combination of cosmological parameters is of the sort \Omega \h_{50}^{\mu} n^{\nu}=0.75+/-0.25, where \mu=1.3 and \nu=3.7,2.0 for \Lambda CDM models with and without tensor fluctuations respectively. For open CDM, without tensor fluctuations, the powers are \mu=0.9 and \nu=1.4.
astro-ph
astro-ph
LARGE-SCALE MASS POWER SPECTRUM FROM PECULIAR VELOCITIES Racah Institute of Physics, The Hebrew University, Jerusalem 91904, Israel I. ZEHAVI This is a brief progress report on a long-term collaborative project to measure the power spectrum (PS) of mass density fluctuations from the Mark III and the SFI catalogs of peculiar velocities.1 2 The PS is estimated by applying maximum likelihood analysis, using generalized CDM models with and without COBE normalization. The applica- tion to both catalogs yields fairly similar results for the PS, and the robust results are presented. , 1 Introduction In the standard picture of cosmology, structure evolved from small density fluctua- tions that grew by gravitational instability. These initial fluctuations are assumed to have a Gaussian distribution characterized by the PS. On large scales, the fluc- tuations are linear even at late times and still governed by the initial PS. The PS is thus a useful statistic for large-scale structure, providing constraints on cosmol- ogy and theories of structure formation. In recent years, the galaxy PS has been estimated from several redshift surveys.3 In this work, we develop and apply like- lihood analysis4 in order to estimate the mass PS from peculiar velocity catalogs. Two such catalogs are used. One is the Mark III catalog of peculiar velocities,5 a compilation of several data sets, consisting of roughly 3000 spiral and elliptical galaxies within a volume of ∼ 80 h−1Mpc around the local group, grouped into ∼ 1200 objects. The other is the recently completed SFI catalog,6 a homogeneously selected sample of ∼ 1300 spiral field galaxies, which complies with well-defined criteria. It is interesting to compare the results of the two catalogs, especially in view of apparent discrepancies in the appearance of the velocity fields.7, 8 2 Method 8 9 9 1 l u J 9 1 v 2 9 0 7 0 8 9 / h p - o r t s a : v i X r a Given a data set d, the goal is to estimate the most likely model m. Invoking a Bayesian approach, this can be turned to maximizing the likelihood function L ≡ P(dm), the probability of the data given the model, as a function of the model parameters. Under the assumption that both the underlying velocities and the observational errors are Gaussian random fields, the likelihood function can be written as L = [(2π)N det(R)]−1/2 exp(cid:16)− 1 i=1 is the set of observed peculiar velocities and R is their correlation matrix. R involves the theoretical correlation, calculated in linear theory for each assumed cosmological model, and the estimated covariance of the errors. ij dj(cid:17) , where {di}N 2 PN i,j diR−1 The likelihood analysis is performed by choosing some parametric functional form for the PS. For each assumed PS, one can calculate the likelihood function and going over the parameter space find the the PS parameters that provide the 1 Figure 1: Likelihood analysis results for the flat ΛCDM model with h = 0.6. lnL contours in the Ω − n plane are shown for SFI (left panel) and Mark III (middle). The best-fit parameters are marked by 's' and 'm' on both, for SFI and Mark III respectively. The right panel shows the corresponding PS for the SFI case (solid line) and for Mark III (dashed). The shaded region is the SFI 90% confidence region. The three dots are the PS calculated from Mark III by Kolatt and Dekel (1997),10 together with their 1σ error-bar. maximum likelihood. Confidence levels are estimated by approximating −2lnL as a χ2 distribution with respect to the model parameters. Note that this method, based on peculiar velocities, essentially measures f (Ω)2P (k) and not the mass density PS by itself. Careful testing of the method was done using realistic mock catalogs,9 designed to mimic in detail the real catalogs. We use several models for the PS. One of these is the so-called Γ model, where we vary the amplitude and the shape-parameter Γ. The main analysis is done with a suit of generalized CDM models, normalized by the COBE 4-yr data. These include open models, flat models with a cosmological constant and tilted models with or without a tensor component. The free parameters are then the density parameter Ω, the Hubble parameter h and the power index n. The recovered PS is sensitive to the assumed observational errors, that go as well into R. We extend the method such that also the magnitude of these errors is determined by the likelihood analysis, by adding free parameters that govern a global change of the assumed errors, in addition to modeling the PS. We find, for both catalogs, a good agreement with the original error estimates, thus allowing for a more reliable recovery of the PS. 3 Results Figure 1 shows, as a typical example, the results for the flat ΛCDM family of models, with a tensor component in the initial fluctuations, when setting h = 0.6 and varying Ω and n. The left panel shows the lnL contours for the SFI catalog and the middle panel the results for Mark III. As can be seen from the elongated contours, what is determined well is not a specific point but a high likelihood ridge, constraining a degenerate combination of the parameters of the form Ω n3.7 = 0.59 ± 0.08, in this case. The right panel shows the corresponding maximum-likelihood PS for the two catalogs, where the shaded region represents the 90% confidence region obtained from the SFI high-likelihood ridge. These results are representative for all other PS models we tried. For each 2 catalog, the different models yield similar best-fit PS, falling well within each oth- ers formal uncertainties and agreeing especially well on intermediate scales (k ∼ 0.1 h Mpc−1). The similarity, seen in the figure, of the PS obtained from SFI to that of Mark III is illustrative for the other models as well. This indicates that the peculiar velocities measured by the two data sets, with their respective error estimates, are consistent with arising from the same underlying mass density PS. Note also the agreement with an independent measure of the PS from the Mark III catalog, using the smoothed density field recovered by POTENT (the three dots).10 The robust result, for both catalogs and all models, is a relatively high PS, with P (k)Ω1.2 = (4.5 ± 2.0) × 103 (h−1Mpc)3 at k = 0.1 h Mpc−1. An extrapolation to smaller scales using the different CDM models gives σ8Ω0.6 = 0.85 ± 0.2. The error- bars are crude, reflecting the 90% formal likelihood uncertainty for each model, the variance among different models and between catalogs. The general constraint of µ nν = 0.75 ± 0.25, where µ = 1.3 and the high likelihood ridges is of the sort Ω h50 ν = 3.7, 2.0 for ΛCDM models with and without tensor fluctuations respectively. For open CDM, without tensor fluctuations, the powers are µ = 0.9 and ν = 1.4. For the span of models checked, the PS peak is in the range 0.02 ≤ k ≤ 0.06 h Mpc−1. The shape parameter of the Γ model is only weakly constrained to Γ = 0.4 ± 0.2. We caution, however, that these results are as yet preliminary, and might depend on the accuracy of the error estimates and on the exact impact of non-linearities.2 Acknowledgments I thank my close collaborators in this work A. Dekel, W. Freudling, Y. Hoffman and S. Zaroubi. In particular, I thank my collaborators from the SFI collaboration, L.N. da Costa, W. Freudling, R. Giovanelli, M. Haynes, S. Salzer and G. Wegner, for the permission to present these preliminary results in advance of publication. References 1. S. Zaroubi, I. Zehavi, A. Dekel, Y. Hoffman and T. Kolatt, ApJ 486, 21 (1997). 2. W. Freudling, I. Zehavi, L.N. da Costa, A. Dekel, A. Eldar, R. Giovanelli, M.P. Haynes, J.J. Salzer, G. Wegner, and S. Zaroubi, ApJ submitted (1998). 3. M.A. Strauss and J.A. Willick, Phys. Rep. 261, 271 (1995). 4. N. Kaiser, MNRAS 231, 149 (1988). 5. J.A. Willick, S. Courteau, S.M. Faber, D. Burstein and A. Dekel, ApJ 446, 12 (1995); J.A. Willick, S. Courteau, S.M. Faber, D. Burstein, A. Dekel and T. Kolatt, ApJ 457, 460 (1996); J.A. Willick, S. Courteau, S.M. Faber, D. Burstein, A. Dekel and M.A. Strauss, ApJS 109, 333 (1997). 6. R. Giovanelli, M.P. Haynes, L.N. da Costa, W. Freudling, J.J. Salzer and G. Wegner, in preparation. 7. L.N. da Costa, W. Freudling, G. Wegner, R. Giovanelli, M.P. Haynes and J.J. Salzer, ApJ 468, L5 (1996). 8. L.N. da Costa, A. Nusser, W. Freudling, R. Giovanelli, M.P. Haynes, J.J. Salzer and G. Wegner, MNRAS submitted (1997). 3 9. T. Kolatt, A. Dekel, G. Ganon and J. Willick, ApJ 458, 419 (1996). 10. T. Kolatt and A. Dekel, ApJ 479, 592 (1997). 4
astro-ph/0501041
1
0501
2005-01-04T14:23:35
A unified description of anti-dynamo conditions for incompressible flows
[ "astro-ph", "physics.flu-dyn" ]
A general type of mathematical argument is described, which applies to all the cases in which dynamo maintenance of a steady magnetic field by motion in a uniform density is known to be impossible. Previous work has demonstrated that magnetic field decay is unavoidable under conditions of axisymmetry and in spherical or planar incompressible flows. These known results are encompassed by a calculation for flows described in terms of a generalized poloidal-toroidal representation of the magnetic field with respect to an arbitrary two dimensional surface. We show that when the velocity field is two dimensional, the dynamo growth, if any, that results, is linear in one of the projections of the field while the other projections remain constant. We also obtain criteria for the existence of and classification into two and three dimensional velocity results which are satisfied by a restricted set of geometries. In addition, we discuss the forms of spatial variation of the density and the resistivity that are allowed so that field decay still occurs for this set of geometries.
astro-ph
astro-ph
A unified description of anti-dynamo conditions for incompressible flows A. Mangalam Indian Institute of Astrophysics, Koramangala Bangalore 560034, INDIA Internet: [email protected] A general type of mathematical argument is described, which applies to all the cases in which dynamo maintenance of a steady magnetic field by motion in a uniform density is known to be im- possible. Previous work has demonstrated that magnetic field decay is unavoidable under conditions of axisymmetry and in spherical or planar incompressible flows. These known results are encom- passed by a calculation for flows described in terms of a generalized poloidal-toroidal representation of the magnetic field with respect to an arbitrary two dimensional surface. We show that when the velocity field is two dimensional, the dynamo growth, if any, that results, is linear in one of the projections of the field while the other projections remain constant. We also obtain criteria for the existence of and classification into two and three dimensional velocity results which are satisfied by a restricted set of geometries. In addition, we discuss the forms of spatial variation of the density and the resistivity that are allowed so that field decay still occurs for this set of geometries. 1. Introduction The steady state dynamo problem in a uniform density can be stated as follows. Given a uniform electrically conducting fluid, contained in a volume, what condition of must be satisfied by the velocity v of a steady motion in order that a steady field B can be maintained by dynamo interaction between the motion and the field? This classical problem of the magnetic dynamo concerns the question of the amplification or maintenance of the magnetic field in cases where the induction equation is valid. The equations yield only decaying solutions when the velocity and magnetic fields are both axisymmetric [1, 2] or if the geometry has planar symmetry [3]. In both situations, the velocity and magnetic field can be three dimensional but do not depend on at least one of the coordinates. In situations where the velocity is two dimensional, but the magnetic field is three dimensional, the impossibility of dynamo action has been proven if the flow is planar [4, 5, 6] or spherical [7, 8]. In the following, we use a generalized toroidal-poloidal representation of the magnetic field, B, with respect to an arbitrary two dimensional surface and derive two scalar equations for the poloidal and the toroidal potentials from the induction equation. We also prove a result, which is an extension of the one in [9], that incompressible two-dimensional velocity flows in situations other than in the above mentioned antidynamo theorems (where the field decays), lead to linear growth in one of the field components and are otherwise slow. Here, the fluid velocity is two dimensional in the sense that it lies entirely on surfaces which can be described by χ(r) = constant. The approach taken here also lends itself to a unified and simpler exposition of the previously cited results. In addition, we consider the special cases of spatially variable forms of the density and magnetic diffusivity. 1 A closer study of the cases in which the dynamo maintenance of a steady field is impossible may throw light on the general dynamo problem. The results suggest that the number if cases in which dynamo maintenance is impossible is restricted. We derive such criteria for the existence of antidynamo result for a given geometry. 2. Normal projection of the induction equation The starting point of all the above investigations is the well known induction equation in MHD ∂tB + ∇ × (B × v) = −∇ × (η∇ × B), with the constraint ∇ · B = 0, (1) (2) where η = c2/4πσ is the magnetic viscosity and σ is the conductivity. As in the above cases, the fluid is assumed to be incompressible (∇ · v = 0). We make the additional simplification of taking η to be uniform. Later, we discuss the effects of relaxing the latter simplification. The RHS of the equation (1) represents resistive dissipation whereas the LHS contains a term which represents stirring of the B field by fluid motions. In analogy with the heat conduction equation, this shows that in a static fluid the fields decay, while stirring may induce field generation. Consider the fluid velocity and the magnetic field to be described in terms of components perpendicular and parallel to the surfaces defined by χ(r) = constant. The quantities, Bχ ≡ B · ∇χ and vχ ≡ v · ∇χ, satisfy the reduced form of the induction equation projected normal to the surface dtBχ − B · ∇vχ − η(∇2B)χ = 0, (3) where dt ≡ ∂t + v · ∇, and the constraint (2) and incompressibility condition was taken into account. One can expand the third term of (3) using identities (35) -- (37), and obtain dtBχ = η(∇2Bχ − ∇ · [(∇χ · ∇)B + (B · ∇)∇χ]) + B · ∇vχ, (4) where the last term, B · ∇vχ, equals zero when the velocity fields lie on the surfaces, χ(r) = const. Now employing the identity (36) on the LHS, writing the first term on the RHS in terms of ∇ · (Bχ∇Bχ) after multiplying throughout by Bχ, and integrating, while taking vχ = 0 (we relax this later), this can be further reduced to the dissipation theorem [9], 1 2 dtZV B2 χd3r = −ηZV {(∇Bχ)2 + BχΘ(B, χ)}d3r, (5) where a surface integral over ∇·(Bχ∇Bχ) obtained from Gauss's theorem vanishes at large distances. Here we have introduced a useful quantity Θ(Y, χ) ≡ ∇ · [(∇χ · ∇)Y + (Y · ∇)∇χ] = [2∂k(Yi) + Yk∂i]∂i∂kχ + ∇χ · ∇(∇ · Y). (6) It can be seen that the last term in equation (6) vanishes if Y is solenoidal or if ∇ · Y is independent of a coordinate directed along ∇χ. The only positive contribution, leading to growth, can come from the tensor term in Θ on the RHS of (6). Clearly, if the surface is planar (χ(r) = z) or spher- ical (χ(r) = r2/2) the term becomes zero (∂i∂k equals 0 or δik, respectively, and in the latter case the condition, (2) needs to be further applied), implying that Bχ decays. This point was made by [9, 10]. Hereafter, we suppress the notation Θ(Y, χ) to Θ(Y) unless χ is specified. In the following, 2 where we keep the treatment general (by keeping the velocity and magnetic fields three dimensional), we find the same term occuring in the surface projection of the induction equation. This enables us to expand upon these conditions for field decay and generalize the antidynamo results cited earlier. 3. Parallel projection of the induction equation It is convenient to express the magnetic field in terms of the local "poloidal" and "toroidal" compo- nents B = BP + BT = ∇ × ∇ × (ψ∇χ) + ∇ × (Φ∇χ), (7) where ψ and Φ are the generalized poloidal and toroidal flux functions. This is analogous to the description of magnetic field given by [11, 4] for spherical geometry. We express the field in a coordinate system in which a special coordinate, q, is normal to the surfaces and is given by χ(r) = f (q). (8) The connection to the corresponding formulae in spherical geometry lies in the fact that any smooth surface has a local radius of curvature. Therefore, we implicity demand that the surface be smooth (have first derivatives defined). Next, using properties (39, 34) it can be seen that Bχ = ∇ × (∇ψ × ∇χ) · ∇χ = ∇ · [(∇ψ × ∇χ) × ∇χ] = ∇ · [(∇ψ · ∇χ)∇χ − (∇χ)2∇ψ] = −(∇χ × ∇)2ψ ≡ −∇2 kψ. (9) In order to examine the local components parallel to the surface, we take a normal projection of the curl of the induction equation (1), ∂t[(∇ × B) · ∇χ)] + ∇χ · [∇ × ∇ × (B × v)] = −∇χ · [∇ × ∇ × (η∇ × B)], (10) After expanding B, the operand of the time derivative in the resulting equation can be reduced in a fashion similar to Eq. (9) Defining, C ≡ ∇ × B, and using Eqs. (35) and (11), the term on the RHS of Eq. (10) yields, (∇ × B) · ∇χ = −∇2 kΦ + Θ(∇ × [ψ∇χ]). −η(curl3B) · ∇χ = η(∇2C) · ∇χ = η(cid:0)∇2Cχ − ∇ · [(∇χ · ∇)C + (C · ∇)∇χ](cid:1) kΦ − ηΘ(C) + η∇2Θ(∇ × (ψ∇χ)), = −η∇2∇2 (11) (12) after performing manipulations identical to those required in obtaining Eq. (6). The second term of equation (10) after applying Eq. (39), is −∇χ · {∇ × ∇ × [v × (∇ × ∇ × (ψ∇χ)) + v × (∇Φ × ∇χ)]}. (13) The second term may be evaluated in steps as follows v × (∇Φ × ∇χ) = vχ∇Φ − (v · ∇Φ)∇χ ∇ × [v × (∇Φ × ∇χ)] = ∇vχ × ∇Φ − ∇(v · ∇Φ) × ∇χ −∇χ · ∇ × {∇ × [v × (∇Φ × ∇χ)]} = −∇ · [∇Φ(∇χ · ∇vχ) − ∇vχ(∇χ · ∇Φ)] −∇2 k(v · ∇Φ). (14) 3 Defining, D ≡ v × BP, Dχ via (A.1) reduces to Dχ = v · (BP × ∇χ). The first term in Eq. (13) will reduce to −∇χ · [∇ × ∇ × D] = −Θ(D) + ∇2Dχ where the properties in the calculation (12) are used. 4. Results (15) (16) Now one can write the equation describing the evolution of the parallel components of the field by including all the terms simplified to the forms given in Eqs. (9) -- (15), and rearranging terms, as ∇2 k(dtΦ) − η∇2∇2 kΦ − ηΘ(C) = ∇ · [∇vχ(∇χ · ∇Φ) − ∇Φ(∇χ · ∇vχ)] The full form of the corresponding equation for the normal component is [dt − η∇2]∇2 kψ − ηΘ(B) = −(B · ∇)vχ, (18) +(∂t − η∇2)Θ(∇ × [ψ∇χ]) − Θ(D) + ∇2Dχ. (17) where the second term involves the normal component of velocity. 4.1 Non-diffusive flows (η = 0) An exclusion theorem proved by [9] that states that two dimensional non-diffusive flows with a stationary velocity field and the property v · a = 0 (the flux helicity density) under the gauge condition, ∇ · a = 0, where a is the vector potential for the velocity, will lead to a conservation of B · a. We sketch their proof below. We write the Euler equation as where all potential forces are collected in w. We can then write, under a gauge condition ∇ · a = 0, ∂tv = f ; f = ∇w − (v · ∇)v, (19) dt(a · v) = v · ∇(a · v) + 2a · f + ∇ · (a × curl−1f ). Next we substitute for f and use −2a((v · ∇)v) = ∇ · (av2) + 2(v × ω) · a enroute to obtain dt(v · a) = 2a · (v × ω) + ∇ · [(v · a)v + a × curl−1f + 2(w − v2/2)a]. (20) (21) (22) This implies that the flux helicity of the flow lines, Hv ≡ R v · a d3x, is conserved for Beltrami flows (v ∝ ω), or for potential flows (ω = 0). Similarly, in the flux freezing limit of the induction equation, ∂tB = ∇ × (v × B); ∂ta = curl−1f we can obtain the following after some transformations dt(a · B) = B · ∇(v · a) + B · curl−1f . (23) (24) 4 Now it is easy to see that for stationary flows (f = 0), if the flux helicity, Hv, is zero everywhere, that the cross helicity, Hc ≡ R B · a d3x, is steady. We now generalize this result by not imposing any limitation on the gauge transformation- as it is impossible to always simultaneously satisfy the conditions v · a = 0 and ∇ · a = 0. When the velocity is two dimensional (vχ = 0) and η → 0, it is clear from (18) that Bχ is an integral of motion and can- not grow with time. Further, if one takes a stationary velocity field with v · a = 0 and ∇ · a = 0, this implies that the flow lines are along the intersection of two surfaces; i.e. v = ∇ξ × ∇χ. As a result, Bξ is another constant of motion (vξ = 0). So the induction, v × B = Bξ∇ξ − Bχ∇χ, is independent of time and the growth of the remaining component of B, (along v), can utmost be linear. Hence, we can conclude the intuitive result that flows with zero linkage (and ∇ · a = 0) cannot lead to an exponential growth of the field, but utmost to a linear growth of the field in the direction of the flow. 4.2 Antidynamo theorems Based on the equations (18) and (17), we can immediately divide the antidynamo results for two and three dimensional velocity fields. For the 3D results, we take both the velocity and the magnetic field to be three dimensional but invariant along the special coordinate q (∂q = 0). In the case of 2D results, the velocity field is two dimensional, vχ = 0, and the magnetic field is three dimensional. Under either of these conditions the first term on the RHS of (17) vanishes. The advection-diffusion operator (dt − η∇2) can only manipulate the field and cannot cause growth (cf. (5)). Here, an antidynamo case is defined as a situation in which the flux functions, ψ and Φ → 0 everywhere (as t → ∞) with the boundary conditions that these flux functions vanish at remote surfaces which enclose the volume of fluid. So the strategy in finding antidynamo situations is to identify the conditions when one equation completely decouples from the other and the decay of the corresponding flux function kills the source terms in the other. It is natural to consider spherical and planar geometries first, since Θ(B) and Θ(C) would then be zero (cf. (6)). It is to be noted that Θ(∇ × [ψ∇χ]) is zero for spherical, planar and axisymmetry. 4.3 The cases of spherical and planar geometries For spherical geometry (χ = 1 2 r2) or planar geometry (χ = z), the equations (17) -- (18) reduce to [dt − η∇2]Φ = Q(ψ, v) [dt − η∇2]∇2 kψ = −B · ∇vχ where Q(ψ, v) ≡ [∇2 k] −1 {∇2Dχ − Θ(D) − ∇ · [∇Φ(∇χ · ∇vχ)]}, (25) (26) (27) k represents L2, the angular momentum operator in the spherical case or ∂ 2 and ∇2 y in the planar case. In the two dimensional case (vχ = 0), we have a source term only in the toroidal equation, (25), while ψ decays. According to the definition of D, as ψ → 0, D → 0 and the RHS of (27) vanishes. Therefore Q → 0 and Φ will decay when t → ∞. This follows from the arguments after (5), and from potential theory which demands that the mean value of [∇2 (0) is zero in the volume enclosed by a surface on which it vanishes. Physically, the normal component diffuses out and the field is confined to two dimensions; as a result, the field is transported like a scalar, and hence decays. x + ∂ 2 k]−1 Now, in the three dimensional planar case (χ = z, ∂z = 0) there is a source term only in the poloidal equation, (26). It can be easily seen from (6) and (15), that the toroidal source terms involving D and ∇χ · ∇ = ∂z are zero. Subsequently, as Φ → 0, B · ∇ → Bz∂z(= 0) and ψ decays. 5 It is interesting that in the three dimensional spherical case none of the source terms in the above pair of equations are zero to begin with, and hence dynamo action occurs. 4.4 The case of axisymmetry In axisymmetry (χ = φ, ∇χ · ∇ = −2∂φ = 0) and D is poloidal. As a result, Q = 0, as the term involving D is zero [cf. (16)], while Θ(A, φ) = −(2/)∂(Aφ/), expressed in the cylindrical coordinates, (, φ, z). Then the equations, (17) -- (18), simplify to D2[dt − ηD2]Φ = 0, (28) [dt − η−2D22](Bφ/) = −B · ∇(vφ/), ∂, and is known as the Stokes operator, and ∇2 (29) k represents [−1 φ × ∇]2 = where D2 ≡ ∇2 − 2 −2D2. Since the D2 term can be reduced to a divergence term and subsequently to a surface integral by Gauss's theorem which vanishes due to the dipole behavior of the field at large radius, it does not contribute to growth; see for example, [4](p. 114). Therefore Φ → 0, B · ∇ → Bφ∂φ(= 0) and ψ decays according to similar arguments in [4] ( p. 115). 4.5 Effects of spatial variation of density and resistivity It is easy to see from the form of the continuity equation in steady state or under the anelastic approximation when the density is dependent only on the special coordinate q, ρ∇ · v + vq∇qρ = 0, (30) that ∇ · v = 0 is valid when vq = 0. Therefore only the above 2D results still hold. Non-uniform resistivity introduces a term (∇η × ∇χ) · (∇ × B) in the equation (18) for ψ. This term is zero if η is a function only of q and hence does not alter the decay of ψ. This was commented upon by [6] for the cases of spherical and planar geometries. We now show that Φ also decays in some special cases. The spatial variation of η in q, however introduces a nonvanishing term (∂zη)(∂zΦ) for planar geometry or (1/r)(∂rη)(∂r[Φr]) for spherical geometry in equation (25). The 3D planar case follows trivially. Now when vχ = 0, one can invoke a theorem on the resulting elliptic equation [12, 5], which states that only the constant solution (Φ ≡ 0) is possible under the condition of Φ vanishing at large distances. Similarly for axisymmetry, η(φ) introduces the term (∂φη)C in (28). When there is no differential rotation, Bφ → 0 and C which depends solely on Bφ, vanishes and Φ decays as before. Therefore, the above 3D planar and 2D planar, spherical, and axisymmetric results are still valid if η is a function only of q. Also, if there is no differential rotation, axisymmetric fields cannot be maintained if η depends on φ. It shown that in axisymmetry [13], poloidal fields cannot be maintained even by a compressible fluid with η as a function of space and time. 5. Concluding remarks In this paper we have cast the induction equation (17) -- (18) in a geometry given by the surfaces χ(r) = constant. This was useful in extending a previous result for incompressible two dimensional flows while unifying, classifying, and simplifying the proofs of the known antidyanamo results, and thereby providing some new insights into the structure of the induction equation. In order to deduce the general conditions of decay, we can consider the order of decay of the flux functions, ψ and Φ. 6 If Φ is to decay first, then all the potential source terms which are on the RHS of (17) containing ψ and D should be zero. This includes the condition that ∇2Dχ − Θ(D) = 0. (31) The above equation automatically ensures that Θ(∇ × [ψ∇χ]) = ∇χ · [∇ × ∇ × ∇ × (ψ∇χ)] is zero (c.f . (16)). The equation, (31) is true for planar and axisymmetry. Consequently, the 3D results of planar and axisymmetry follow, as the first term on the RHS of (17) which involves ∇χ · ∇ is zero. On the other hand, if ψ is to decay first then the RHS of (18) should be zero demanding the 2D condition, vχ = 0. Further, ZV [Θ(B) − ∇2Bχ]d3r ≥ 0, (32) must hold to ensure that ψ decays as per (5) and the boundary conditions that the flux functions vanish at the remote boundaries. This takes care of the decay of Φ since the remaining source terms, which involve ψ, in (17) vanish. The condition, (32), is true for planar, spherical and axisymmetric k]−1(0) geometries. The unifying aspect of treatment of the boundary conditions used here is that [∇2 is zero with vanishing flux at the boundaries. We have been able show the impossibility of dynamo action (as defined in §4) with η = const and ∇ · v = 0, exists only for a restricted group of geometries as allowed by (31) and (32) for the case of 3D and 2D results respectively. In addition, the above results need further qualifications that v is bounded for all time and that "spiky" time dependent behavior is excluded [14]. As suggested in a classic paper [15], using a different approach, that the impossibility of a steady dynamo for a given geometry depends on the existence of an arbitrary current, j′, such that R j · j′d3r = 0, and that only a restricted set satisfies this equation. Here we have derived specific conditions that determine these geometries. However, a more rigorous analysis of the above two conditions is needed to find the set of all possible χ that satisfies the above criteria. In a paper, in preparation, I investigate the geometries that satisfy the above criteria. Acknowledgment: I thank P. H. Roberts for his valuable comments and for a critical reading of the manuscript. Appendix: Formulae referenced in the text a · (b × c) = b · (c × a) = c · (a × b) a × (b × c) = b(c · a) − c(a · b) ∇ × (∇ × a) = ∇(∇ · a) − ∇2a ∇(a · b) = (a · ∇)b + (b · ∇)a + a × (∇ × b) + b × (∇ × a) ∇ · (a × b) = b · (∇ × a) − a · (∇ × b) ∇ × (a × b) = a(∇ · b) − b(∇ · a) + (b · ∇)a − (a · ∇)b ∇{×, ·}(ψa) = ∇ψ{×, ·}a + ψ∇{×, ·}a (33) (34) (35) (36) (37) (38) (39) References 1. Cowling T G, Mon. Not. R. Astron. Soc., 94 (1934), 39. 7 2. Backus G E, Chandarsekhar S, Proc. Nat. Acad. Sci., 42 (1956), 105. 3. Zeldovich Ya B, J. Expt. Theo. Phys., 31 (1956), 154. 4. Moffatt H K, Magnetic Field Generation in Electrically Conducting Fluids, Cambridge, Cambridge University Press (1978). 5. Lortz D, Phys. Fluids, 11 (1968), 913. 6. Zeldovich Ya B, Ruzmaikin A A, J. Expt. Theo. Phys., 78 (1980), 980. 7. Bullard E C, Gellman H, Phil. Trans. R. Soc. Lond., A247 (1954), 213. 8. Backus G E, Ann. Phys., 4 (1958), 372. 9. Ruzmaikin A A, Sokoloff D D, Geophys. and Astrophys. Fluid Dynamics, 16 (1980), 73. 10. Zeldovich Ya B, Ruzmaikin A A, Sokoloff D D, Magnetic Fields in Astrophysics, New York, Gordon & Breach (1983). 11. Chandrasekhar S, Hydrodynamic and Hydromagnetic Stability, Dover, New York (1961). 12. Vekua I N, Generalized Analytic Functions, Oxford, Pergamon Press (1962). 13. Hide R, Palmer T N, Geophys. and Astrophys. Fluid Dynamics, 19 (1982), 301. 14. James R W, Roberts P H, Winch D E, Geophys. and Astrophys. Fluid Dynamics, 15 (1980), 149. 15. Cowling T G, Q. J. Mech. Appl. Math., 10 (1957), 129. 8
astro-ph/0103277
1
0103
2001-03-17T13:24:18
The Unique Signature of Shell Curvature in Gamma-Ray Bursts
[ "astro-ph" ]
As a result of spherical kinematics, temporal evolution of received gamma-ray emission should demonstrate signatures of curvature from the emitting shell. Specifically, the shape of the pulse decay must bear a strict dependence on the degree of curvature of the gamma-ray emitting surface. We compare the spectral evolution of the decay of individual GRB pulses to the evolution as expected from curvature. In particular, we examine the relationship between photon flux intensity (I) and the peak of the \nu F\nu distribution (E_{peak}) as predicted by colliding shells. Kinematics necessitate that E_{peak} demonstrate a power-law relationship with I described roughly as: I=E_{peak}^{(1-\zeta)} where \zeta represents a weighted average of the low and high energy spectral indices. Data analyses of 24 BATSE gamma-ray burst pulses provide evidence that there exists a robust relationship between E_{peak} and I in the decay phase. Simulation results, however, show that a sizable fraction of observed pulses evolve faster than kinematics allow. Regardless of kinematic parameters, we found that the existence of curvature demands that the I - E_{peak} function decay be defined by \sim (1-\zeta). Efforts were employed to break this curvature dependency within simulations through a number of scenarios such as anisotropic emission (jets) with angular dependencies, thickness values for the colliding shells, and various cooling mechanisms. Of these, the only method successful in dominating curvature effects was a slow cooling model. As a result, GRB models must confront the fact that observed pulses do not evolve in the manner which curvature demands.
astro-ph
astro-ph
The Unique Signature of Shell Curvature in Gamma-Ray Bursts Alicia Margarita Soderberg1 and Edward E. Fenimore2 1 DAMTP, Silver Street, Cambridge CB3 9EW, ENGLAND 2 Los Alamos National Laboratory, Los Alamos NM 87545, USA Abstract. As a result of spherical kinematics, temporal evolution of received gamma- ray emission should demonstrate signatures of curvature from the emitting shell. Specif- ically, the shape of the pulse decay must bear a strict dependence on the degree of curvature of the gamma-ray emitting surface. We compare the spectral evolution of the decay of individual GRB pulses to the evolution as expected from curvature. In particular, we examine the relationship between photon flux intensity (I) and the peak of the νF ν distribution (Epeak) as predicted by colliding shells. Kinematics necessi- tate that Epeak demonstrate a power-law relationship with I described roughly as: I = E(1−ζ) peak where ζ represents a weighted average of the low and high energy spectral indices. Data analyses of 24 observed gamma-ray burst pulses provide evidence that there exists a robust relationship between Epeak and I in the decay phase. Simulation results, however, show that a sizable fraction of observed pulses evolve faster than kinematics allow. Regardless of kinematic parameters, we found that the existence of curvature demands that the I − Epeak function decay be defined by ∼ (1 − ζ). Ef- forts were employed to break this curvature dependency within simulations through a number of scenarios such as anisotropic emission (jets) with angular dependencies, thickness values for the colliding shells, and various cooling mechanisms. Of these, the only method successful in dominating curvature effects was a slow cooling model. As a result, GRB models must confront the fact that observed pulses do not evolve in the manner which curvature demands. 1 Introduction to the Kinematic Model The simulated pulses described in this study were created through code based strictly on kinematics. Simulated shells were collided with one another, thereby conserving energy and momentum and the resulting energy was distributed into standard Band function spectra. Isotropic emission from the merged shell was (initially) assumed where the entirety of the shell is modeled to be gamma-ray active. Figure 1 demonstrates the geometry of the model. Time of arrival is determined by the angle, θ, at which the emitting patch lies with respect to the line of sight. Off axis emission is received later than on axis emission by a factor of R(1 − cosθ). The emitting shell has a slight thickness defined between R/c = tmax and R/c = t0. Photons within shell volume dV contribute to the pulse shape between received times, T and T + dT . As a result of the relativistic motion of the shell, the volume of emitting material which contributes to the received signal at any time is constant. Emitting patches on the shell which fall between the two ellipsoids labeled T and T +dT will arrive at the detector within this range of received time. 2 Soderberg & Fenimore t o t max T T+dT dV R θ R (1.0 - cos θ) Fig. 1. Geometry of the Kinematic Model 2 Discussion: The Robust Curvature Dependency Results of the kinematic studies demonstrate that the I − Epeak relationship is a robust indicator of shell curvature. The strength of this relation was analyzed by varying both Band and kinematic parameters for the shell model. Observed pulses, however, do not demonstrate this dependence (see Figures 2 and 3). In the attempt to break spherical symmetry and reduce the dependence of pulse shape on curvature effects, more complicated emission models were simulated. These models enabled further testing to examine the possibility of additional depen- dencies which were not included in the original kinematic code. Models explored jetting the model emission into an opening angle between 0.1-5.0 degrees and allowing for intrinsic angular dependencies of the Lorentz factor and/or Epeak across the skullcap. Off-axis shell collisions were simulated such that the colli- sion time was not instantaneous in the rest frame of the central engine and the initial photon emission occurred at an angle outside the critical beaming angle. Various thicknesses were applied to the emitting shell but this only proved to distort the rise time of the pulse and did not have any effect on the shape of the pulse decay or the I − Epeak relationship. Models also explored fast and slow Fig. 2. Comparison with BATSE pulses. Simulations were compared with a data set of 24 pulses selected from the BATSE GRB Spectral Catalog I (Preece et al.,1999). The top figure displays the robustness of the expected and observed I − Epeak relationships for GRB921207. The solid line represents the expected decay index as predicted by colliding shells. It is evident that there is a large discrepancy between the data and the simulations. The same is true for the bottom figure which displays the results for GRB970201.Through these cases, the severity of this discrepancy can be clearly seen. The Unique Signature of Shell Curvature in Gamma-Ray Bursts 3 cooling mechanisms. Generally, it was found that the curvature dependence is fairly difficult to break, and requires either grossly distorted geometries and/or relatively long cooling time scales. It was found that the emitting shell must Fig. 3. Comparison of Expected I − Epeak Decay with Observed I − Epeak Decay from BATSE Pulses. The solid line represents the I − Epeak decay index demanded by colliding shells. Curvature and special relativity impose such a relationship because later portions of the pulse arrive (1−ζ) from off-axis emission. This results in an expected decay index of I = E peak . The majority of BATSE observations lie below the solid line therefore indicating that observed I − Epeak decay is slower than the decay predicted by kinematics. cool for a time period of approximately tcool = R/c in the detector rest frame (where R/c < Γ 2). This corresponds to a comoving time of t′ cool = R/cΓ and an arrival time of Tcool = R/cΓ 2 by the standard transformations. As a result, the observed cooling time was comparable in length to the duration due to cur- vature (e.g. 104 s). Such slow cooling overwhelmed the curvature dependency with cooling effects throughout the entire length of the pulse, thereby allowing for a new pulse shape evolution. It is emphasized that slow cooling was the only method included in this study which was able to break the robust curvature de- pendency as imposed by the kinematics of two colliding shells. Typical cooling times, however, are commonly quoted as being shorter than the duration of the pulse (e.g. < 104 s). A remedy to this situation is to minimize the timescale on which curvature effects can be detected by reducing the radius and/or increasing the bulk Lorentz factor of the emitting shell. This, in turn, allows for a relatively shorter cooling time. Long cooling times, however, face a number of problems including efficiency considerations which must be addressed. References 1. Band, D. et al.: ApJ 413, 281 (1993) 2. Crider, A., et al.: ApJ 519, 206 (1999) 3. Fenimore, E. E., Madras, C. D., & Nayakshin, S.: ApJ 473, 998 (1996) 4. Preece, R. D., et al.: ApJ Supp. 126, 19 (1999) 5. Rybicki, G. B. & Lightman, A. P.: Radiative Processes in Astrophysics, 1979 6. Sari, R., & Piran, T.: ApJ 485, 270 (1997) 7. Summer, M. C., & Fenimore, E. E.: AIP Proc. 428, 765 (1998)
astro-ph/0406135
1
0406
2004-06-05T14:56:47
Sub-Galactic Clumps at High Redshift: A Fragmentation Origin?
[ "astro-ph" ]
We investigate the origin of the clumpy structures observed at high redshift, like the chain galaxies. We use a three dimensional chemodynamical simulation describing the dynamics of stars and a two-phase interstellar medium, as well as feedback processes from the stars. For high efficiency of energy dissipation in the cold cloud medium, the initially gaseous disk fragments and develops several massive clumps of gas and stars. We follow the evolution of the individual clumps and determine their masses, metallicities and velocities. A few dynamical times after fragmentation of the disk, the clumps merge to build a massive bulge. Calculating HST- and UBVRIJHKLM-colors, including absorption by interstellar dust, we determine the morphologies and colors of this model in HST images. Several peculiar morphological structures seen in the HDF can be well-explained by a fragmented galactic disk model, including chain galaxies and objects consisting of several nearby knots.
astro-ph
astro-ph
Sub-Galactic Clumps at High Redshift: A Fragmentation Origin? Andreas Immeli1, Markus Samland1, Pieter Westera2, Ortwin Gerhard1 ABSTRACT We investigate the origin of the clumpy structures observed at high redshift, like the chain galaxies. We use a three dimensional chemodynamical simulation describing the dynamics of stars and a two-phase interstellar medium, as well as feedback processes from the stars. For high efficiency of energy dissipation in the cold cloud medium, the initially gaseous disk fragments and develops several mas- sive clumps of gas and stars. We follow the evolution of the individual clumps and determine their masses, metallicities and velocities. A few dynamical times after fragmentation of the disk, the clumps merge to build a massive bulge. Calculat- ing HST- and UBVRIJHKLM-colors, including absorption by interstellar dust, we determine the morphologies and colors of this model in HST images. Several peculiar morphological structures seen in the HDF can be well-explained by a fragmented galactic disk model, including chain galaxies and objects consisting of several nearby knots. Subject headings: Galaxies: evolution -- Galaxies: formation -- Galaxies: high- redshift -- Galaxies: structure -- Galaxies: peculiar 1. Introduction In the redshift range 0.5 < z < 3, galaxies evidence a large diversity of morphological types (e.g. Abraham et al. 2001; Steidel et al. 1996; van den Bergh et al. 1996, vdB96). Al- though some of the unusual morphological structures can be explained by the morphological K-correction, NICMOS observations (Dickinson 2000) show that many galaxies indeed have a rest frame optical morphology that cannot be attached to the traditional Hubble scheme. Examples are the so-called chain galaxies that show elongated knotty structures (Cowie et al. 1995, CHS95). Different scenarios have been proposed to explain these structures. 1Astronomisches Institut der Universitat Basel, Venusstrasse 7, CH-4102 Binningen, Switzerland 2Observat´orio do Valongo, Universidade Federal do Rio de Janeiro, Ladeira do Pedro Antonio, 43, CEP 20080-090, Rio de Janeiro, Brazil -- 2 -- CHS95 suggested that chains lie in the redshift range 0.5 − 3 and have a mass comparable to that of a present-day galaxy. They speculate that these objects may be linear arrangements in space where star formation, once turned on, triggers further star formation along the line of maximum density. In some models the chains form in colliding supershells blown out of massive starburst galaxies (Taniguchi & Shioya 2001). Dalcanton & Shectman (1996) argued that LSB galaxies are local counterparts to chains at high or intermediate redshift. O'Neil et al. (2000) suggested, from a comparison with less inclined objects, that the chains do not belong to a new galaxy class but are knotty disk like structures seen edge on. This is consistent with recent observations from (Elmegreen et al. 2004a,b), who find that the clump colors in face-on clumpy objects are similar to the colors of clumps in chain galaxies. Here we investigate a model of a gaseous disk that becomes unstable and develops several clumps of gas and stars. The model presented here describes one of the evolutionary paths that a disk can take, in the sequence investigated in Immeli et al. (2004), with a higher- resolution simulation. The evolution of this disk is similar as in the scenario proposed by Noguchi (1998), but our model for the star-forming two-phase interstellar medium, including feedback processes, allows us to keep track of stellar ages and metallicities, and thus to determine realistic luminosities and colors for a direct comparison with observations in the Hubble Deep Field (HDF). We show here that several of the unusual morphological types in the HDF are well-described by a fragmented disk model seen from different viewing angles. 2. The Model We use a two-phase model for the interstellar medium, consisting of a hot, low-density phase and a cold cloud medium from which stars are formed. We describe this system with a three-dimensional chemodynamical evolution code, which combines a hydrodynamical grid code for the two phases of the interstellar medium (ISM) with a particle mesh code for the stars. The interactions between the different ISM phases are described in Samland & Gerhard (2003, SG03). For the star formation rate (SFR) we use a Schmidt Law (Schmidt 1959), ρsf = csf · ρα cld with α = 1.5 and csf consistent with the star formation (SF) rule derived by Kennicutt (1998). The energy released from supernovae mostly goes into heating the hot phase, but also heats the cold phase. Most of the kinetic energy of the cloud fluid is in the motions of single clouds relative to the bulk flow. This kinetic energy can be dissipated by inelastic collisions (Larson 1969) and augmented by supernova feedback (McKee & Ostriker 1977). Its energy dissipation rate, here described by the parameter ηc, is not well-determined, and may well vary between galaxies. One expects that it depends on the geometrical structure of the -- 3 -- clouds, on whether a major part of the dense medium is arranged in filaments, and on their self-gravitating structure and magnetic fields (Kim et al. 2001; Balsara et al. 2001). The influence of ηc on the dynamical evolution of gas-rich disks is investigated in more detail in Immeli et al. (2004); together with the infall rate it determines the amount of star-forming cold gas in the disk. Here we compare one of their fragmenting disk models (ηc = 0.5) with observations of high-redshift galaxies. The setup of our model describes an early and rapid formation of a massive galactic disk in a static dark halo. According to Sommer-Larsen et al. (2002) the delayed infall of the baryonic matter into the relaxed halo can solve the angular momentum problem arising in ΛCDM structure formation simulations. The primordial gas enters the simulation volume at z = 7 kpc vertically and uniformly distributed over a radius of 17 kpc, with a rotation velocity equal to the circular velocity at the infall point, and infall velocity 20 km/s. The infall rate is 120 M⊙yr−1 during one Gyr, resulting in a total baryonic mass of 1.2 · 1011 M⊙. The simulation volume has a diameter of 37.2 kpc and a vertical height of 14 kpc with a spatial resolution of 300 pc in the horizontal and 120 pc in the vertical direction. We have also done the simulation at lower resolution, with similar results, indicating that the outcome is not sensitive to the resolution used (Immeli et al. 2004). The chemodynamical model provides ages and metallicities of the stars formed, as well as ISM densities and metallicities. This enables us to calculate colors of the model at different redshifts, including absorption, using the method of Westera et al. (2002) except that we adopt here a three times lower absorption coefficient. 3. Results and Comparison to Observations 3.1. Global Evolution The infall of the baryons into the halo leads to the build-up of a star-forming gaseous disk. Fig. 1 shows the resulting evolution of the star formation rate (SFR) in the model. The energy input from the supernovae type II dominates that from the infall during most of the evolution and prevents the rapid formation of a massive disk on a free fall timescale. However, at around 700 Myr the gas disk becomes unstable on large scales and begins to fragment. The lower-mass stellar system follows the gravitational potential perturbations induced by the gas. Fig. 2 shows a face-on view of the model evolution from 500-1500 Myr in terms of observed HST F606W surface brightness. The model was shifted to the redshift range indicated in the frames. -- 4 -- Fig. 1. -- Star formation rate of the gas-rich disk model discussed in this paper. -- 5 -- To quantify the fragmentation we have used the asymmetry parameter A (Abraham et al. 1996) in rest frame U-band. The evolution of A is very similar to that of the SFR, which illustrates that the SF is driven by fragmentation. The high symmetry in the first two images reflects the symmetric infall of the gas. The pressure from the SF in the disk and the pressure from the infalling material causes the development of the ring-like structure at the border of the stellar disk, visible in the second image. This ring structure represents less than 10% of the total mass of the cloudy medium in the disk at this time. Yet the enhanced SF in this structure, due to feedback-induced large density fluctuations, causes a very prominent UV emission shifted to F606W at the redshift considered. It is important to note that the disk shows its clumpy structure also in the H-band (Fig. 3), which would be traced by NICMOS observations. This emphasizes the fact that the clumps are not only regions of high star formation in an underlying smooth disk, but that the disk itself is fragmented, forming stars vigorously in several dense clumps. This makes it hard to distinguish such clumpy disks from a merger event on the basis of observed surface brightness maps alone. The maximum SFR is reached in this simulation during the main fragmentation phase, after about 1 Gyr. The morphology of the disk at this time is bracketed by panels 3 and 4 of Fig. 2. The SFR at this time is around 220 M⊙yr−1, corresponding to a strong starburst galaxy. Indeed, many high-redshift objects may be starburst galaxies. Lowenthal et al. (1997) report similarities in stellar emission and interstellar absorption lines between z ∼ 3 galaxies and local starburst galaxies. The color selection criteria for Lyman Break Galaxies (LBG) also strongly favor starburst galaxies (Steidel et al. 1996). Age determinations of stellar populations and enhanced abundances of α-elements in LBGs (Carollo & Lilly 2001) indicate that the very high SFRs in these high-redshift objects persist only for a few hundred Myrs. The clumps that form in the disk during the fragmentation phase spiral to the center, building a massive bulge (last two panels of Fig. 2). The clumps lose their angular momen- tum by dynamical friction against the massive arms they generate in the disk. Because a substantial fraction of the mass in the region of interest by then consists of baryonic mate- rial, the timescale for this spiral-in phase is relatively short, of the order of two disk rotation times. This is typical for all models with a fragmenting gas disk discussed in Immeli et al. (2004). The strongly asymmetric potential leads to a redistribution of angular momentum, which is partially carried away by stars which leave the simulation volume. The high mass of the bulge is explained by the efficiency of the angular momentum transfer during the fragmen- tation phase. -- 6 -- Without newly infalling material there will be no major changes in the global structures of the galaxy after 2.5 Gyr. Formation of a bar is prevented due to the high-mass bulge. Numerical investigations of Noguchi (1999) showed a qualitatively similar evolution of a gaseous disk. Two main differences in his work are that the dissipative evolution is described by only a single sticky particle phase, and that in his model the high SF threshold allows SF only after fragmentation of the disk. The stability properties of multi-phase star-gas disks such as the model analyzed here, and the implications for bar and bulge formation, are discussed in more detail in Immeli et al. (2004). 3.2. Comparison with Observations In Fig. 4 the model morphology at different times during the fragmentation phase is compared to the morphologies of some objects in the HDF. Clearly, several HDF morpholo- gies can be well explained by the fragmented disk model. In the first row of Fig. 4 we show a phase of enhanced spiral arms at about 1.35 Gyr in the model, induced by the merging of the last two massive clumps, and a similar object observed in the HDF. In the second row we compare the model at 1.3 Gyr to a clumpy structure at high redshift. Seen edge-on (third row), the model resembles a chain galaxy during its fragmentation phase. CHS95 reported observations of chain galaxies (chains), a new population of high red- shift galaxies observed with HST in the Hawaii Survey Fields, with high major-to-minor axis ratios and very blue colors. vdB96 also found chain galaxies in the HDF. As already mentioned in the introduction, there are different explanation scenarios for the chain galaxies. Recent observations point to the fact that chain galaxies are indeed edge-on disks (Elmegreen et al. 2004a). Comparison of the clump colors with the colors of clumpy face on disks (sometimes called clump clusters) leads to similar results, indicating that these clump clusters are indeed face-on counterparts of chain galaxies. Additionally these authors found that the distribution of axial ratios for chain galaxies and clumpy disks is similar to the distribution of local disk galaxies. Given the small number of chains compared to approximately 1000 known high redshift galaxies (Giavalisco 2002), it is likely that these objects are in a short evolutionary phase. Our clumpy disk model indicates that the interpretation of O'Neil et al. (2000) and Elmegreen et al. (2004a) is correct. It shows chain structures when viewed edge-on and during a period -- 7 -- Fig. 2. -- Fragmentation phase of the star-forming disk model, shown in observed F606W surface brightness, starting at 0.5 Gyr at top left and continuing in 200 Myr intervals. For K-correction and surface brightness dimming the middle right panel was shifted to z = 1.5, a typical redshift for chain galaxies. Redshifts of other panels are relative to the z = 1.5 panel and are indicated in each map. The frames are 40 kpc a side. HST resolution and a detection limit of 28.21 mag (Williams et al. 1996) were used. Angular diameters were calculated using a ΛCDM cosmology with ΩM = 0.3, ΩΛ = 0.7, h = 0.7. -- 8 -- Fig. 3. -- Fragmentation phase of star-forming disk model, as in Fig. 2, but in NICMOS H160 surface brightness. Here we used a limiting surface brightness of 25.05, corresponding to the 10σ limit of 26.1 mag in an aperture of 0."7 diameter in the observations of Ellis et al. (2001). Notice that the clumpy disk structure is also visible in infrared passbands, showing that not only the light distribution, but also the mass distribution of the disk is fragmented. -- 9 -- Fig. 4. -- Comparison of observed HST F606W surface brightness of the star-forming disk model (left panels) with observations from vdB96 (right panels). For this comparison the model was shifted to the indicated photometric redshifts of the observed galaxies (Fernandez- Soto et al. 1999). -- 10 -- of very high SFR; see Fig. 5. Because this period is short compared to a Hubble time, these objects will be relatively rare, and because of the high SFR, they are very blue. The model therefore naturally explains also the observation of CHS95 that a large fraction of chains is very blue. A comparison of the model colors with those of CHS95 is shown in Fig. 6. Best agreement is obtained if our model is shifted to redshifts between 0.8 and 1.8 (see also Immeli et al. (2003)). Comparing the color profile of the chain galaxies with observations (Fig. 7) shows also good agreement. The model reproduces the flat profile, which in our model is a direct consequence of the constant-surface density infall. The typical exponential profile observed in present-day disk galaxies only emerges in the instability phase. The scale length of the resulting exponential disk depends on the infall radius. Because of the strong instabilities in the disk, gas and stars are dynamically heated. The gas, on the other hand, is also cooled by dissipative cloud collisions. In the present model, the velocity dispersion of the gas takes values between 30 and 50 km/s, about 25% of the maximal rotation velocity. This is consistent with estimates from observations (Elmegreen et al. 2004a). Also the number of clumps and their masses are in perfect agreement with these observations. 3.3. Local galaxies Do gas-rich disk systems at low redshift also tend to form fragments with enhanced SF? An example of a nearby galaxy in which the fragmentation process may be taking place, is the gas-rich, blue starburst galaxy NGC 7673 (Homeier & Gallagher 1999). This object has a remarkably clumpy morphological appearance, seen in both the R-band and Hα, even though the Hα velocity field is that of a regular, rotating disk. Dwarf irregular galaxies observed in the local universe generally have a high gas fraction and often a disk-like structure. Recent observations (Billett et al. 2002) show that there is a tendency for star formation to be concentrated in localized regions of high column density. Due to the lower mass of the dwarf galaxies, one cannot directly compare them to our model. In particular, the dynamical influence of the dark halo is likely to be more important in dwarf galaxies; explicit models of such lower mass galaxies are needed. -- 11 -- 4. Merger or Fragmentation? Kinematical data will be important to further clarify the nature of chain galaxies. While in our fragmented disk model the massive clumps should still show the underlying disk rotation, in a merger scenario no similar alignment of the clump velocities is expected. Fig. 8 shows the predicted influence of the clumps on the rotation curve of the gaseous disk; deviations from the smooth rotation profile are up to 100 km/s. These deviations are due to the gravitational influence of the massive clumps; they are much larger than expected from the velocity dispersion in the disk alone. In summary, while the basic rotation pattern remains visible during the fragmentation phase, it is severely disturbed by the innermost brightest knots. Additionally, the metallicities of the clumps in a merger event are expected to vary significantly, depending on the mass and evolution history of the merging clumps. Contrary- wise, one expects similar metallicities for the clumps in a fragmented disk. We investigate the metallicity of five clumps selected in the fragmented disk model as indicated in Fig. 9. We get abundance differences of up to 0.25 dex (Table 1), depending on the masses of the clumps (∼ few 109 M⊙ for those in Fig. 9). No differences in oxygen-to-iron ratios can be measured, due to dominance and young age of the starburst. Many authors report observations of multiple knots (Driver et al. 1995; Steidel et al. 1996, vdB96). The fragmentation scenario naturally explains objects consisting of several clumps, whereas in a merging scenario it is much less likely to see more than three nearby clumps merging at the same time. Also, the synchronized colors often observed in these clumpy objects (Abraham et al. 2001) are naturally explained with the fragmented disk scenario. In the present model, the mean age of the stars seen edge-on in the clumps is constant within 70 Myr. At least some of the observed multi-clump systems may therefore represent fragmented disks. This is confirmed by the recent observations of Elmegreen et al. (2004a): a merger event would lead to more spheroidal systems than the thin chain galaxies. 5. Conclusions A short formation timescale of a galactic disk due to a high dissipation rate for the cold gas phase leads to fragmentation and to the formation of a clumpy disk with an enhanced SFR in the clumps. After the fragmentation phase the clumps fall to the center building a massive bulge. Subsequent bar formation is prevented by the massive bulge. Chain and multi-clump morphological structures, as well as synchronized colors observed in high redshift objects, can be well explained by a fragmented disk in a gas rich, single galaxy. -- 12 -- Chain, double and tadpole galaxies may be different evolutionary states of a fragmented disk. Our model suggests that these galaxies are in their formation process and are observed during their relatively short fragmentation phase, with a high SFR, comparable to the model SFR of up to 220 M⊙ yr−1. This high SFR is generated by a disk instability alone, and there is no need for external triggering through interactions or a merger with other galaxies. The effects of single clumps on the mass-weighted rotation curve in our model can be as high as 100 km/s. Nonetheless the underlying rotation signature should be observable. Metallicity differences in the clumps of the fragmented disk are no larger than 0.25 dex, while their mean stellar ages are highly synchronized. Observations to test these predictions are highly desirable. We thank the Schweizerischen Nationalfonds for financial support of this work and the Centro Svizzero di Calcolo Scientifico (CSCS) for giving us the opportunity to use their computing facilities. -- 13 -- Fig. 5. -- The chain galaxy model (at time 1.15 Gyr, shifted to z = 1.6, as in Fig. 4), edge-on and face-on in observed F606W surface brightness. Seen face-on, the chain model resembles the clumpy disks observed by Elmegreen et al. (2004a). -- 14 -- Fig. 6. -- Colors of the model seen edge-on at 1.15 Gyr, shifted to the indicated redshifts (squares). Observational data for chains from CHS95 (stars). -- 15 -- Fig. 7. -- Color profile of the clumpy disk model seen edge-on, after 1.15 Gyr, when shifted to z = 1.6 as in Fig. 4. The positions of the main clumps are indicated by the vertical dashed lines. The figure shows that the typical flat profile for chain galaxies (Elmegreen et al. 2004b) is nicely reproduced, and that the clumps as regions of intense star formation are both bright and blue in V-I color, again as observed. -- 16 -- Fig. 8. -- Mass-weighted rotation curve of the gas after 1.15 Gyr, in the disk plane where the clumps dominate (solid line), and 0.5 kpc above the plane (dashed line). The deviations from the smooth rotation curve are significantly larger than expected purely from the velocity dispersion in the disk. They reflect the gravity of the massive clumps in the disk. -- 17 -- Fig. 9. -- Smoothed stellar mass surface density of the model at 1.2 Gyr with numbering of the clumps. -- 18 -- Clump Mass [109M⊙] 1 2 3 4 5 7.96 4.10 2.02 2.39 0.64 [O/H] -0.32 -0.38 -0.44 -0.42 -0.59 [Fe/H] -0.72 -0.77 -0.84 -0.82 -0.98 [O/Fe] 0.40 0.39 0.40 0.40 0.39 Table 1: Masses and metallicities of the stars in the clumps at 1.2 Gyr (see Fig. 9). Metallicity differences are up to 0.25 dex. [O/Fe] is constant because SNe Ia with a typical time delay of around 1 Gyr have not yet contributed significantly to the chemical abundances. -- 19 -- REFERENCES Abraham, R. G., Tanvir, N. R., Santiago, B. X., Ellis, R. S., Glazebrook, K. & van den Bergh, S. 1996, MNRAS, 279, L47 Abraham, R. G., & van den Bergh, S. 2001, Science, 293, 1273 Balsara, D., Ward-Thompson, D., Crutcher, R. M. 2001, MNRAS, 327, 715 Billett, O. H., Hunter, D. A., & Elmegreen, B. G. 2002, AJ, 123, 1454 Carollo, C. M. & Lilly, S. J. 2001, ApJ, 548, L153 Cowie, L. L., Hu, E. M. & Songaila, A. 1995, AJ, 110, 1576 Dalcanton, J. J., & Shectman, S. A. 1996, ApJ, 465, L9 Dickinson, M. 2000, in "Building Galaxies: From the Primordial Universe to the Present", XIXth Moriond Astrophysics Meeting, ed. F. Hammer et al. (Paris: Ed. Frontieres), 257 Driver, S. P., Windhorst, R. A., & Griffiths, R. E. 1995, ApJ, 453, 48 Ellis, R.S., Abraham, R.G., & Dickinson, M. 2001, ApJ, 551, 111 Elmegreen, D. M., Elmegreen, Bruce G., Hirst, A. C. 2004, ApJ, 604, 21L Elmegreen, D. M., Elmegreen, Bruce G., Sheets, C. M. 2004, ApJ, 603, 74 Fernandez-Soto, A., Lanzetta, K. M., & Yahil, A. 1999, ApJ, 513, 34 Giavalisco, M. 2002, ARA&A, 40, 579 Homeier N. L., Gallagher J. S., 1999, ApJ, 522, 199 Immeli, A., Samland, M., Gerhard, O. 2003, EAS Publications Series 10, "Galactic and Stellar Dynamics", ed. C. M. Boily, P. Pastsis, S. Portegies Zwart, R. Spurzem and C. Theis, pp. 199 Immeli, A., Samland, M., & Gerhard, O. E., Westera P. 2004, A&A, 413, 547 Kennicutt, R. C., 1998 ApJ, 498, 541 Kim, J., Balsara, D., Mac Low, M.-M. 2001, JKAS, 34, 333 Larson, R.B., 1969, MNRAS, 145, 405 -- 20 -- Lowenthal, J. D., Koo, D. C. Guzman, J. G., Phillips, A. C., Faber, S. M., Vogt, N. P., Illingworth, G. D., & Gronwall, C. 1997, ApJ, 481, 673 McKee, C. F., Ostriker, J. P., 1977, ApJ, 218, 148 Noguchi, M. 1998, Nature, 392, 253 Noguchi, M. 1999, ApJ, 514, 77 O'Neil, K., Bothun, G. D., & Impey, C. D. 2000, ApJS, 128, 99 Samland, M., & Gerhard, O. 2003, A&A, 399, 961 Schmidt, M. 1959, ApJ, 129, 243 Sommer-Larsen, J., Gotz, M., & Portinari, L. 2002, astro-ph/0204366 Steidel, C., Giavalisco, M., Dickinson, M., & Adelberger, K. L. 1996, AJ, 112, 352 Taniguchi, Y, & Shioya, Y 2001, ApJ, 547, 146 van den Bergh, S., Abraham, R. G., Ellis, R. S., Tanvir, N. R., Santiago, B. X., & Glazebrook, K. G. 1996, AJ, 112, 359 Westera, P., Samland, M., Gerhard, O., & Buser, R. 2002, A&A, 389, 761 Williams, R. E., et al. 1996, AJ, 112, 1335 This preprint was prepared with the AAS LATEX macros v5.2.
astro-ph/0408295
1
0408
2004-08-16T21:41:52
The SMART Data Analysis Package for the Infrared Spectrograph on the Spitzer Space Telescope
[ "astro-ph" ]
SMART is a software package written in IDL to reduce and analyze Spitzer data from all four modules of the Infrared Spectrograph, including the peak-up arrays. The software is designed to make full use of the ancillary files generated in the Spitzer Science Center pipeline so that it can either remove or flag artifacts and corrupted data and maximize the signal-to-noise in the extraction routines. It may be run in both interactive and batch mode. The software and Users Guide will be available for public release in December 2004. We briefly describe some of the main features of SMART including: visualization tools for assessing the data quality, basic arithmetic operations for either 2-d images or 1-d spectra, extraction of both point and extended sources and a suite of spectral analysis tools.
astro-ph
astro-ph
Draft version November 26, 2011 Preprint typeset using LATEX style emulateap j v. 6/22/04 4 0 0 2 g u A 6 1 1 v 5 9 2 8 0 4 0 / h p - o r t s a : v i X r a THE SMART DATA ANALYSIS PACKAGE FOR THE INFRARED SPECTROGRAPH1 ON THE SPITZER SPACE TELESCOPE 2 S. J. U. Higdon3 , D. Devost3 , J. L. Higdon3 , B. R. Brandl4 , J. R. Houck3 P. Hall3 , D. Barry3 , V. Charmandaris3,5 , J. D. T. Smith6 , G. C. Sloan3 , & J. Green7 Draft version November 26, 2011 ABSTRACT SMART is a software package written in IDL to reduce and analyze Spitzer data from all four modules of the Infrared Spectrograph, including the peak-up arrays. The software is designed to make full use of the ancillary files generated in the Spitzer Science Center pipeline so that it can either remove or flag artifacts and corrupted data and maximize the signal-to-noise in the extraction routines. It may be run in both interactive and batch mode. The software and Users Guide will be available for public release in December 2004. We briefly describe some of the main features of SMART including: visualization tools for assessing the data quality, basic arithmetic operations for either 2-d images or 1-d spectra, extraction of both point and extended sources and a suite of spectral analysis tools. Subject headings: methods: data analysis — techniques: spectroscopic — telescopes: Spitzer Space Telescope 1. INTRODUCTION The Spectroscopy Modeling Analysis and Reduction Tool (SMART) is a software package written in IDL 8 for the analysis of data acquired with the Infrared Spectrograph1 (IRS) on the Spitzer Space Telescope 2 (Werner et al. 2004). The code has been developed for the Unix/Linux operating systems. The IRS comprises four separate spectrograph modules covering the wave- length range from 5.3 to 38 µm with spectral resolutions, R = λ/∆λ ∼ 90 and 600. The modules are named af- ter their wavelength coverage and resolution as Short- Low (SL), Short-High (SH), Long-Low (LL) and Long- High(LH). The SL includes two peak-up imaging cam- eras that have band-passes centered at 16 µm (“blue”) and 22 µm (“red”). For details of the IRS instrument see Houck et al. (2004) and chapter 7 of the Spitzer Ob- servers Manual9 (SOM7). SMART has been designed specifically for IRS data and in particular to extract spectra from observations of faint or extended sources. It has been written with an understanding of both the available Spitzer IRS observ- ing modes and a knowledge of how the contents of various files generated by the Spitzer Science Center (SSC) IRS pipeline can be used to maximize the signal-to-noise in 1 The IRS was a collaborative venture between Cornell Univer- sity and Ball Aerospace Corporation funded by NASA through the Jet Propulsion Laboratory and the Ames Research Center. 2 The Spitzer Space Telescope is operated by JPL, California Institute of Technology for the National Aeronautics and Space Administration. 3 Astronomy Department, Cornell University, Ithaca, NY 14853; [email protected] 4 Leiden Observatory, 2300 RA Leiden, The Netherlands 5 Chercheur Associ´e, Observatoire de Paris, F-75014, Paris, France 6 Steward Observatory, University of Arizona, Tucson, Arizona, 85721 7 Department of Physics and Astronomy, University of Rochester, Rochester, NY 14627 Electronic address: [email protected] 8 The Interactive Data Language, Research Systems, Inc. 9 http://ssc.spitzer.caltech.edu/documents/som/ the extracted spectrum. These three design factors make it a comprehensive and powerful software package for the extraction and analysis of IRS data. SMART is primarily intended to operate on the ba- sic calibrated data (BCD, see SOM7) delivered by the SSC pipeline, but will also operate on the browse quality data (BQD, including both images and wavelength and flux calibrated spectral tables) and 2-d data products from intermediate stages of the SSC pipeline, for exam- ple, the un-flatfielded data. SMART aims to provide the routines necessary for the processing and scientific anal- ysis of IRS data. The main goal is to simplify the tasks of visualizing, organizing, optimally combining and ex- tracting data. The result of this processing are fully flux and wavelength calibrated spectra. Further analysis is available within SMART. Additionally, the spectra can be easily exported (in either FITS, ascii or IDL save set format) to other analysis packages written, for example, in IDL or IRAF. SMART includes software developed by two of the Spitzer Legacy teams. The Molecular Cores to Planet- Forming Disks (C2D) team has developed a code to re- move fringes caused by interference in the detector sub- strate material (Lahuis & Boogert 2002). This software is an enhanced version of the code developed for the In- frared Space Observatory (ISO) Short Wavelength Spec- trometer (Kester et al. 2003). The Formation and Evo- lution of Planetary Systems (FEPS) team have adapted the IDP3-NICMOS package (Schneider & Stobie 2002) to analyze image data from the IRS peak-up cam- The spectral analysis code is based on the eras. (ISAP10 inherited ISO Spectral Analysis Package )(Sturm et al. 1998). The ISAP software is available at http://www.ipac.caltech.edu/iso/isap/isap.html introduc- present The an as is paper tion SMART. A SMART web site at to 10 The ISO Spectral Analysis Package (ISAP) is a joint develop- ment by the LWS and SWS Instrument Teams and Data Centers. Contributing institutes are CESR, IAS, IPAC, MPE, RAL and SRON. 2 Higdon et al. as serves http://isc.astro.cornell.edu/smart/ the repository for the full listing of all functions avail- able in SMART and details of the algorithms used. This website includes a comprehensive SMART Users Guide (SUG) and a set of data reduction recipes aimed at the new user. Each recipe outlines the steps required to produce wavelength and flux calibrated spectra. Both the website and the software will be publicly available in December 2004. In the following section we introduce the main graphical user interfaces (GUIs) used for the interactive analysis of IRS data and briefly describe the experienced user and batch-mode capabilities. 2. SMART GUIS Before starting a SMART analysis session the ob- servers need to fetch their IRS data from the Spitzer archive. SMART is designed to operate on the SSC pipeline basic calibrated data (BCD) FITS files. A BCD is a calibrated, flatfielded 2-d image. The observer ob- tains a BCD image for each IRS exposure (for observ- ing mode and pipeline details, see SOM7). In addition to the BCD file SMART also needs two associated files for each exposure. These files are the uncertainty data and the bad pixel mask. The bad pixel mask has a 16 bit integer assigned to each pixel. Each bit corresponds to a given warning/error condition detected during the pipeline processing. For example a pixel may have suf- fered a cosmic ray hit or may be saturated. A perl script searches the local directory for these 3 files and builds a new FITS file (‘*bcd3p.fits’) for each exposure. Each new FITS file contains the data plane and two extensions: the uncertainty plane and the bad pixel mask plane. 2.1. Project Manager SSC pipeline products are read into the pro ject man- ager and either proceed directly or via image analysis into the ISAP-based Data Evaluation and Analysis GUI (IDEA). Figure 1 presents a flow chart outlining the main graphical user interfaces (GUIs) available in SMART. At first glance some of the GUIs may appear complex, but we remind the reader that the SUG will be available at the SMART website. Figures 2a & 2b show the pro ject manager and dataset GUIs. These form the base for launching different appli- cations and storing the resulting data products. The pro ject manager allows the Spitzer observer to load files by browsing a local directory containing the 2-d images and spectral table files from the archive or import files from an existing local data base. It is designed to han- dle large data sets by grouping them into sets called “pro jects”. BCDs within one pro ject may or may not come from the same IRS module or astronomical target. For example, consider the simple case of a low-resolution observation of a point source covering the full wavelength range (5.3 − 38 µm). This requires the use of both the SL and LL modules. Each module covers its nominal spectral range in two orders via two sub-slits. The de- fault observing mode will obtain two spectra of the target source per sub-slit resulting in eight separate exposures. The resulting spectra from this observation would con- sist of two sets of four spectra from 5.2 - 8.7 µm (SL2), 7.4 - 14.5 µm (SL1), 14.0 - 21.3 µm (LL2) and 19.5 - 38.0 µm (LL2) of the target source. The IRS low res- olution observations always obtain data simultaneously in the two sub-slits, so there are an additional two sets of four spectra, with the same wavelength coverage as above, of the background sky. Entire pro jects can be saved to disk as IDL save sets, which can be imported into new SMART pro jects. Alternatively individual files can be exported from the pro ject to disk in either FITS or ascii table format. The main applications launched from the pro ject manager are described in the following sections. 2.2. Image Display/Analysis 2.2.1. ATV-IRS We have enhanced a version of the image display pro- gram ATV, to work with our IRS spectral 2-d images. The ATV code was developed to visualize both 2-d and 3-d images (Barth 2001). Figure 3 shows an example of data displayed in ATV-IRS. We have enhanced the code so that an over-plot tracing the curved spectral or- ders and the boundaries of the individual resolution ele- ments can be displayed. The cursor position is reported in terms of pixel position (x,y) and flux (pixel value) as well as sky coordinates (right ascension, declination) and spectral wavelength. This is very useful for assessing whether weak features in a spectrum are emission lines or are caused by cosmic-ray hits to the detectors. The viewer can also display the uncertainty and bad pixel mask planes, returning the same information for the cur- sor position. Additional pixels may be flagged at this stage by editing the bad pixel mask. In addition to dis- playing a single BCD, one can make a movie of a stack of images or make a single mosaiced image. A table con- taining the pixel information can also be inspected. If a stack of images is selected the statistics on the cube of pixel data can be displayed in a table. Images from an external archive can also be added to the pro ject man- ager and viewed in ATV-IRS. 2.2.2. IDP3-IRS The Image Display Paradigm 3 (IDP3) is a sophisti- cated photometry software package. It is written in IDL and is designed for the analysis of the Hubble NICMOS data. The IRS has imaging capabilities provided by the two peak-up cameras, each with an ∼ 1 arcmin2 field of view. The blue IRS peak-up camera fills a gap in the wavelength coverage of the Spitzer imagers at 16 µm. Both the red (22 µm) and blue cameras are used for many observations. Elizabeth Stobie at the University of Arizona provided a modified version of IDP3, known as IDP3-IRS, which is optimized for analyzing sources observed with the IRS peak-up mode. 2.2.3. Quick Look Quick Look takes a BCD file and collapses the spectral orders along the dispersion direction to produce an av- erage intensity profile across the source. This provides a convenient tool to search for either extended emission or weak secondary point sources in the low resolution data, which have slits that are 57 ′′ (SL) and 168 ′′ (LL) in the cross-dispersion direction. Figure 4 shows an exam- ple of a serendipitous detection of a weak source located close to the slit center in LL1. The negative stripes in LL2 are caused by the sky subtraction in LL1 using data which has the target source in LL2, see Section 2.4 for a discussion of sky subtraction. SMART 3 2.2.4. Image Operations The standard image operations - averaging, median fil- tering, division, addition and subtraction - are available in SMART. In addition to operations weighted by the uncertainty data, pixels can be discarded according to their bad pixel mask value. This offers a powerful means for discarding corrupted data and improving the signal to noise. For example, consider the median of ten BCDs. First the pixels flagged by a bad pixel mask value are ex- cluded from the calculation of the median value for each pixel in the 128x128 array. A new bad pixel mask is gen- erated for the median data using the ’OR’ operator on all the bad pixel mask values associated with the data used to estimate the median. Automatic co-adding and dif- ferencing are available for SL and LL data. The images are sorted by slit and position within the slit (i.e., nod position) and then co-added. The co-added “on-source” and “off-source” images may then be differenced. 2.3. Spectral Extraction There are currently four extraction routines available in SMART. All methods use a look-up table supplied by the SSC, which traces the spectral orders on the BCD image. This table is converted into an IDL structure known as the “wavesamp”. The extraction routines use the wavesamp in order to sub-set the relevant group of pixels in each resolution element on the array. An exam- ple of the wavesamp trace is shown in Figure 3, where the curved spectral order has been sub-divided into spectral resolution elements. The curvature results in fractional pixels being assigned to a given resolution element. The value of each fractional pixel is scaled by its geometrical area. The BCD data is in units of electron/sec. Each rou- tine estimates the total number of electron/sec in each individual resolution element. The final step is to apply the flux calibration (i.e., electron/sec to Janskys) and stitch the orders together into a single spectrum, using the pipeline “fluxcon” tables. 2.3.1. Ful l Aperture Extraction This is the standard extraction method for SH and LH data. It can also be used for extended ob jects that fill the SL and LL apertures. The pixel values in each resolu- tion element, as defined by the wavesamp, are summed, while accounting for fractional pixels. Prior to extrac- tion, pixel values set to NaN (Not a Number) in the pipeline or flagged by the bad pixel mask are replaced with an average value, which is estimated from the val- ues of the pixels in the same resolution element as the bad pixel. 2.3.2. Column Extraction For SL and LL observations of point sources the stan- dard extraction method is column extraction. A col- umn of pixels centered on the point source is extracted. Figure 5 is an example of the aperture used for a col- umn extraction in LL2. The column traces the spec- tral order and its width in the cross-dispersion direc- tion is scaled with the instrumental point spread func- tion. The user should over-ride the default width only after careful consultation of the manual and help pages ( see http://isc.astro.cornell.edu/smart/). The user defined width is scaled with wavelength, and re- quires additional “on-the-fly” calibration, see Section 3. The pixels in the column in each resolution element are summed, while accounting for fractional pixels. Prior to extraction pixel values set to NaN or flagged by bad pixel mask values are replaced with an average value, which is estimated from the pixels in the column within the same resolution element as the bad pixel. 2.3.3. Extended Source Extraction A column of pixels centered on the extended source is extracted. The column traces the spectral order and its width is constant with wavelength. Again, the pixels in each resolution element within the column are summed, while accounting for fractional pixels. Pixel values set to NaN or flagged by bad pixel mask values are replaced using the method outlined above for column extraction. 2.3.4. Gaussian Extraction Gaussian extraction should only be used with care as it requires additional on-the-fly calibration, see Section 3. The data from the pixels in each individual resolution element are collapsed in the dispersion direction. The resulting 1-d trace is fit with a Gaussian profile. The Gaussian center and width can be frozen to aid the ex- traction of weak sources. Pixel values set to NaN or flagged by bad pixel mask values are excluded from the fit. 2.4. Sky Subtraction The BCD images include sky emission and possible de- tector artifacts, which can be removed in SMART. The first method is applied before extraction and the remain- ing two methods are akin to removing a baseline before measuring a line flux. If no sky image data are available a zodiacal model can be subtracted from the spectra in IDEA. 2.4.1. Super-Sky Subtraction The sky emission is removed by differencing an “on- source” and “super-sky” image prior to extraction. A super-sky image can be created by co-adding multiple “off-source” (i.e., sky) BCDs together. This can be done in SMART using either the image operations GUI or us- ing one of the available scripts. This method is applicable to all four modules. For low resolution data the “super- sky” image may simply be the median of the “off-source data”, which is acquired as part of the standard staring mode observation. For high resolution data separate sky observations are required. 2.4.2. Single-Sky Subtraction The sky emission is removed during the extraction pro- cess. An “off-source” BCD is used for the sky estimate. This can be done in two ways. The first method calcu- lates the median sky pixel value in each resolution ele- ment of the off-source BCD. In the second method, the pixel values in each resolution element of the off-source BCD are plotted as a function of cross-dispersion dis- tance from the center of the slit. A first-order polyno- mial is fit to the resulting intensity profile to estimate the sky level. For full aperture extraction the sky value is scaled with the area of the resolution element. For 4 Higdon et al. column extraction the sky value is scaled to the area of the column with in a given resolution element. For Gaus- sian extraction the sky value is scaled to the width of the Gaussian. The scaled sky value is then subtracted from each respective summed resolution element, column or integrated Gaussian profile in the source spectrum. 2.4.3. Local-Sky Subtraction The Single-Sky methods are applied to the data. How- ever in this instance the sky is calculated from the same BCD that contains the source data. The pixels in a given resolution element that are not part of the source column or Gaussian are used to estimate the sky value. An ex- ample of a selection of suitable sky regions is shown in Figure 5. 2.5. IDEA IDEA, the ISAP-based Data Evaluation and Analy- sis Program, is a comprehensive 1-D spectral analysis package. The code includes the inherited ISAP soft- ware. ISAP was developed for the analysis of spectral data from ISO (SWS/LWS/PHT-S/Cam-CVF) and pro- vides a wealth of routines embedded in an easy-to-use graphical environment. The ability to analyse ISO spec- tra has been preserved so that direct comparisons may be made between data from the two satellites. Figure 6a shows the IDEA GUI, which has been enhanced to fit the special needs of IRS data. The applications GUI is shown in Figure 6b. Processing routines include shift- ing, zapping low signal-to-noise or corrupted data, de- fringing, re-binning, unit conversion, combining spectra with weighted means or medians, filtering and smooth- ing. Analysis routines include line fitting, line identifica- tion, continuum-fitting, synthetic photometry (including the IRAS, ISO and Spitzer filter profiles), zodiacal light modeling, blackbody fitting, de-reddening and template fitting routines. The spectra can be imported/exported as FITS, IDL save sets or ascii tables. 3. EXPERIENCED USER AND BATCH MODE SMART is supplied with a default set of calibration files from the SSC, which can be inspected in the Cal- ibration GUI. However the experienced user can create her/his own set of calibration files and import them into SMART via the Calibration GUI. The extraction rou- tines can be tailored to specific source profiles to maxi- mize the signal to noise. Intermediate pipeline products, for example, the un-flatfielded data, can be substituted for the BCD image files. This can be beneficial for the extraction of faint sources and sources that have weak features. When un-flatfielded data are used the flux cal- ibration is performed on-the-fly using a default set of calibration sources. Both the target and the flux cali- brator are extracted using the same parameters. The ex- tracted calibration source is then used to flux calibrate the extracted target spectrum. Only an experienced user should over-ride the default calibration file selection. SMART is also designed for efficient batch mode pro- cessing. Many of the GUI functions are available in batch mode and we are developing a suite of scripts for the most commonly used functions. For example, both full aperture, column and Gaussian extraction are available in batch mode. The script includes two sky removal op- tions. The first option is to subtract the sky during ex- traction. Alternatively, the sky can be removed using the super-sky method. The data from a given slit and nod position are co-added and then the co-added “on- source” and “off-source” images are differenced. A sin- gle spectrum is then extracted from the median filtered, sky-subtracted image. 4. SUMMARY currently being tested by the SMART is IRS In De- team and the participating legacy teams. cember 2004 we will have a public release of SMART. The code will be available at our web- site, http://isc.astro.cornell.edu/smart/, where the IRS observer can find detailed instructions for down- loading and installing SMART. The site includes a SMART Users Guide and recipes for reducing IRS data. We plan to update and add new functionality to the code as our understanding of the IRS data analysis evolves. For example, we are currently working on an optimized extraction algorithm for both high and low resolution data. Updates will be posted at the web site. We would like to thank the following people, the IRS team and the SSC for their dedicated work in generating the pipeline processed data and for ongoing calibration work; the ISAP team for allowing us to inherit and mod- ify the ISAP code and the referee, Eckhard Sturm, for his swift endorsement of this paper. This work is based [in part] on observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Tech- nology under NASA contract 1407. Support for this work was provided by NASA through Contract Number 1257184 issued by JPL/Caltech. REFERENCES Barth, A. J. 2001, in ASP Conf. Ser., Vol. 238, Astronomical Data Analysis Software and Systems X, eds. F. R. Harnden, Jr., F. A. Primini,& H. E. Payne (San Francisco: ASP), 385 Houck, J. R., et al. 2004 ApJS, 154 (in press) Lahuis, F. & Boogert, A.SFChem 2002: Chemistry as a Diagnostic of Star Formation, proceedings of a conference held August 21-23, 2002 at University of Waterloo, Waterloo, Ontario, Canada N2L 3G1. Ed Charles L. Curry and Michel Fich. To be published by NRC Press, Ottawa, Canada, 63 Kester, D. J. M., Beintema, D. A. & Lutz, D. 2003, in The the ISO Mission, proceedings of a calibration legacy of conference held Feb 5-9, 2001, ed. L. Metcalfe, A. Salama, S.B. Peschke and M.F. Kessler. Published as ESA Publications Series, ESA SP-481. European Space Agency, 2003, 375 Schneider, G. & Stobie, E., 2002, Astronomical Data Analysis Software and Systems XI ASP Conference Series, Vol 281, 2002 D.A. Bohlender, D. Durand, and T.H. Handley eds ”Pushing the Envelope: Unleashing the Potential of High Contrast Imaging with HST” 2002, 382 - 386 SMART 5 Stored Projects IDL SMART Project format IRS Data FITS & table format Calibration Widget Set calibration files & extraction parameters Image Viewer ATV Image Viewer IDP3 Image Viewer Quickview Project Manager Create, store & recall projects Dataset Manager Load, view & manipulate images Extraction Auto or Manual Manual Extraction Tool Fig. 1.— SMART Flow Chart Processed Data ASCII, FITS, XDR Image Operations Co−add, arithmetic IDEA Window View & Manipulate Spectra IDEA Applications Synthetic photometery, line & template fitting, defringing, etc. Fig. 2a.— The SMART Pro ject Manager GUI Fig. 2b.— The SMART Dataset GUI Sturm, E. et. al. 1998, Astronomical Data Analysis Software and Systems VII, A.S.P. Conference Series, Vol. 145, 1998, eds. R. Albrecht, R.N. Hook and H.A. Bushouse,,1998, 161 Werner, M. W. et al, 2004, ApJS, 154 (in press) 6 Higdon et al. Fig. 3.— The SMART ATV-IRS GUI - Data from IRS LL2 is displayed. Over-plotted are the resolution elements and the column which has been selected for spectral extraction. SMART 7 Fig. 4.— The SMART Quick Look GUI - sky subtracted BCD data from the IRS LL is displayed in the upper image window. LL1 has been collapsed in the dispersion direction and is displayed in the plot window. A serendipitous source is observed close to the centre of the slit. The negative stripes in LL2 are caused by the sky subtraction in LL1 using data which has the target source in LL2. 8 Higdon et al. Fig. 5.— The SMART Mandefine GUI - BCD data from the IRS long low spectrometer is displayed in the upper image window. LL2 has been collapsed in the dispersion direction and is displayed in the plot window. Overlaid on this plot are the column boundaries (dotted-lines) and sky regions (dashed-lines) for column extraction. Fig. 6a.— The SMART IDEA GUI - this is the main GUI for viewing 1-d spectral data SMART 9 Fig. 6b.— The SMART IDEA Applications GUI - The ma jority of the spectral reduction and analysis routines are accessed from this GUI
astro-ph/0105413
2
0105
2001-05-30T21:41:58
Type Ia Supernovae, the Hubble Constant, the Cosmological Constant, and the Age of the Universe
[ "astro-ph" ]
The age of the Universe depends on both the present-day Hubble Constant and on the history of cosmic expansion. For decelerating cosmologies such as Omega_m= 1, the dimensionless product H_0,t_0<1 and modestly high values of the Hubble constant H_0 > 70 would be inconsistent with a cosmic age t_0 larger than 12 Gyr. But if Omega_Lambda > 0, then H_0,t_0 can take on a range of values. Evidence from the Hubble diagram for high redshift Type Ia supernovae favors Omega_Lambda~0.7 and H_0,t_0 ~ 1. Then, if H_0 lies in the range 65--73, the age of the Universe, t_0, is 14+/-1.6 Gyr.
astro-ph
astro-ph
Astrophysical Ages and Time Scales ASP Conference Series, Vol. TBD, 2001 T. von Hippel, N. Manset, C. Simpson Type Ia Supernovae, the Hubble Constant, the Cosmological Constant, and the Age of the Universe John L. Tonry Institute for Astronomy, University of Hawaii The High-Z Supernova Search Team http://cfa-www.harvard.edu/cfa/oir/Research/supernova/HighZ.html Abstract. The age of the Universe depends on both the present-day Hubble Constant and on the history of cosmic expansion. For decelerat- ing cosmologies such as Ωm= 1, the dimensionless product H0 t0 < 1 and modestly high values of the Hubble constant H0 > 70 would be inconsis- tent with a cosmic age t0 larger than 12 Gyr. But if ΩΛ > 0, then H0 t0 can take on a range of values. Evidence from the Hubble diagram for high redshift Type Ia supernovae favors ΩΛ ∼ 0.7 and H0 t0 ∼ 1. Then, if H0 lies in the range 65 -- 73, the age of the Universe, t0, is 14 ± 1.6 Gyr. 1. It has been an interesting five years! Five years ago, the combination of deep seated belief in inflation, implying Ω = 1, and stellar age estimates near 15 Gyr seemed to require H0 ∼ 40. Measurements of H0 ∼ 60 and Ωm ∼ 0.3 in clusters notwithstanding, Bartlett, Blanchard, Silk, and Turner wrote a provocative paper entitled, "The Case for a Hubble Constant of 30 km/s/Mpc." Persuaded by the power of theoretical reasoning, Joe Silk bet Brian Schmidt and me a case of Scotch that H0 < 60. While Joe has not yet paid up, in the past 5 years he has moved closer to the source of Scotch while the Hubble constant has moved to 60 and beyond. The new element is that supernovae have made the connection between Ω = 1 and the cosmic age more flexible because of plausible evidence for cosmic acceleration. A Danish-English team (Norgaard-Neilsen et al. 1989) initiated a program to find supernovae in clusters of galaxies at redshifts of 0.3-0.5, with the idea that they could distinguish the effects of cosmic deceleration, as expected in an Ωm = 1 universe by measuring the peak apparent magnitudes of supernova light curves. The observational problem was to find these faint (m∼21-22) and distant supernovae near the peak of their light curves. But small detectors kept this pioneering effort from yielding significant results. The Supernova Cosmology Project (SCP) based at Lawrence Berkeley Lab forged ahead with further attempts to find distant supernovae by extending the methods of the Danes to bigger, faster telescopes. After abandoning attempts to instrument the AAT prime focus for this purpose, they used the standard large format detectors at the Kitt Peak National Observatory starting in 1992 1 SNIa, H0, Λ, and t0 2 (Perlmutter 1995). By 1997, they had a preliminary result (Perlmutter et al. 1997) based on observations of seven supernovae discovered in 1994 and 1995. By comparing their supernovae with the sample at low redshift, they concluded that the evidence favored a universe with high matter density Ωm = 0.88 ± 0.6. They argued that the supernova data at that point placed the strongest constraint on the possible value of the cosmological constant, with their best estimate being ΩΛ = 0.06. Since 1995, there have been two groups pursuing evidence on cosmic decel- eration using the Hubble diagram for supernovae. Our High-Z Supernova Search Team, steered by Brian Schmidt, and encompassing workers on four continents and one mid-Pacific island, found our first supernova in 1995 (Schmidt et al. 1998). From three additional supernovae studied with HST the High-Z Team found a low value of Ωm, showing that ordinary matter could not close the Uni- verse (Garnavich et al. 1998). Our first results on ΩΛ were reported at the Dark Matter conference in February of 1998 by Filippenko and Riess (1998) and published in the Astronomical Journal in September 1998 (Riess et al. 1998). In the same period, the SCP revised their analysis of earlier data (Perlmutter et al. 1998) and then independently reported evidence on ΩΛ in the Astrophysical Journal in June 1999 (Perlmutter et al. 1999). 2. SNIa constraints on t0 Even this first round of supernova observations, which emphasized a sample near z∼0.5, provided a good constraint on the difference Ωm-ΩΛ which, since Ωm mea- sures deceleration, and ΩΛ measures acceleration, translates into a surprisingly tight constraint on H0 t0. The present samples of SNIa published, in hand, and being reduced by the two teams provide a statistically robust measurement that ΩΛ > 0. The impor- tant questions now are whether the supernovae at large redshift are really the same as the supernovae nearby and whether exotic forms of grey dust might ob- scure both the supernovae and our understanding of cosmology (Aguirre, 1999, 2000). The observational approach to answer these questions is to use spectra to examine the question of homogeneity (Coil et al. 2001) and multicolor ob- servations over a wide range of wavelengths (as might be done with a superb 8-m infrared telescope at the world's best site) to constrain the properties of intergalactic dust (Riess et al. 2000). So far, although the distant supernovae could have failed these tests, they seem indistinguishable from the SNIa nearby. A more ambitious test for the cosmological origin of the observed effect is to extend the data set to higher redshift. While dimming due to evolution or dust would most naturally lead to larger effects at higher redshift, cosmological effects could have the opposite sign due to cosmic deceleration at early epochs (z ∼ 1.5), followed by a transition to acceleration in the more recent past. 2.1. What does a supernova look like? Looking at real data helps develop an understanding of the observational issues in discovering and measuring light curves for high redshift supernovae. Here we illustrate the appearance of a z=0.81 supernova, SN 1999fj, as observed in a series of images from October through December 1999. Large format detectors SNIa, H0, Λ, and t0 3 on telescopes at good seeing sites are the chief requirement for efficient surveys to find type Ia supernovae at m ∼ 24.5. Follow-up observations with 8-m class telescopes enable us to construct light curves that can be used to place each SNIa firmly on the Hubble diagram. Figure 1. SN 1999fj at four epochs. 03 Oct 99 has the supernova barely above our detection threshold using I band at the CFHT. 03 Nov 99 is the discovery epoch. The reader is invited to find the new dot. This image covers approximately 1/3000 the area of the detector array, so the real search is automated. 15 Nov 99 shows an image in 0.5′′ seeing from the VLT, and 15 Dec 99 shows an image in 0.9′′ seeing from the Keck telescope. 2.2. Light curves from Fall 1999 One of the key developments that makes SNIa so useful as standard candles is the discovery by Mark Phillips (Phillips 1993) that the luminosity of a SNIa is related to its rate of decline after maximum light. This approach, refined by Hamuy et al. (1996) and by Riess, Press & Kirshner (1996) allows the distance of a well-observed SNIa with two color data to be determined to better than 10%. SNIa, H0, Λ, and t0 4 The light curves for high redshift supernovae are obtained in filters which can be transformed back to rest frame B and V with good precision. Time di- lation, a signature of cosmic expansion, is a powerful effect for supernovae near z∼ 1, transforming 40 days in the observer's frame to 20 days in the supernova's own rest frame (Leibundgut et al. 1995). Photometric calibration and scrupu- lous subtraction of galaxy light are serious problems for this work which become more difficult at high redshift. Nevertheless, images like those in Figure 1 can be used to construct light curves as shown in Figure 2. Figure 2. Fall 1999, ranging from z=1.05 to z=0.28 A sampling of two-color light curves for 6 supernovae from 3. Meditations on H0 What is the observational evidence on H0? Many techniques have been de- veloped for measuring extragalactic distances and the ones we love the best are SNIa (Jha et al. 1999) and the Surface Brightness Fluctuation method, reviewed recently by Blakeslee, Ajhar, & Tonry (1999). Ajhar et al. 2001 demonstrate that these methods are internally consistent: as Figure 3 shows, distances mea- SNIa, H0, Λ, and t0 5 sured by SNIa and by SBF to the same galaxies are consistent within the quoted errors. This suggests that both methods are in good shape. But what is H0, so diligently sought through the decades? This depends entirely on the sample of Cepheids used and the distances assigned to those Cepheids by expert workers in that field. If you use the distances to galaxies with Cepheids as determined by various papers by the Key Project (Ferrarese et al. 2000, Freedman et al. 2001, Gibson & Stetson 2001) to define the absolute magnitude of SNIa and SBF, you get H0 = 73, 75, and 77 for both SNIa and SBF. If you use the distances to the same galaxies using the same Cepheid observations as reduced by Saha et al (1997) and most recently compiled by Parodi et al. (2000), you get H0 = 65 according to the methodology used by the High-Z team, or H0 = 58.5 according to Parodi et al. Similarly, the distances to the SBF calibrating galaxies depend entirely on the Cepheid distances. The Hubble constant is not determined by SNIa or by SBF alone: these methods give excellent relative distances to galaxies and tie the cosmic expansion firmly to the local calibrators, but the calibration by Cepheids is now the largest uncertainty in measuring the local value of the Hubble Constant. For the rest of this paper, we will either adopt H0 = 73 as a best-guess value, or else a probability distribution which consists of two Gaussians of fractional width 0.1, centered on H0 = 73 and H0 = 65, with a 2/3, 1/3 weighting. It ain't perfect, but it's our gut feeling of where H0 really lies. Figure 3. they get their zero points from the same set of Cepheids. There is no discrepancy between the SNIa and SBF when 4. Differential Hubble diagram for SBF and SNIa The key evidence on a value for H0 t0 comes from Figure 4. All the supernova and SBF distances are consistent at low redshift, and the supernova data indi- cate that a model with ΩΛ=0 does not fit at z∼ 0.5. This is the case for an accelerating universe, whether urged on by a cosmological constant or by some- SNIa, H0, Λ, and t0 6 Figure 4. The cosmological diagram for SBF and supernovae, show- ing observed luminosity distance (units of Glyr) divided by cz and 1 + z/2 (i.e. divided by luminosity distance in an empty universe) as a function of redshift. Both sets of distances are normalized to an H0 = 73 zeropoint; the left hand intercept is therefore H −1 ∼ t0; the right axis is labelled in magnitudes. The darker gray region shows where cosmological models with ΩΛ = 0 lie, and the light and dark gray regions show where flat models (Ω0 = 1) lie. The black down- turning curve shows an (0.3, 0.7) cosmology, enhanced by the darkest gray region of ±500 km/s which corresponds to ±2.5σ thermal pecu- liar velocities. The large black points are medians in various redshift bins, and the error bars are estimates of the uncertainty in the median judged from the scatter of the contributing points. At z ∼ 0.5 these points lie very significantly above the region permitted if ΩΛ = 0, and well away from ΩM = 0.3. The black dashed line illustrates how a sys- tematic error that is proportional to z diverges from agreement with an (0.3, 0.7) cosmology for z > 0.5. 0 thing which varies with time. The best fit model has ΩM = 0.3 and ΩΛ = 0.7. None of the observers is satisfied with the current state of the statistical errors (about twice as large for each SCP supernova as for the High-Z data) or with current limits on possible systematic effects that might make distant supernovae dimmer. No Big Bang 3 2 % 9 . 7 9 0 . 5 - 0 = q 0 0 = q 0 . 5 0 = q g a ti n a ti n r g r e l e e l e c A c c D e Clo s e d O p e n Expands to Infinity ^ Recollapses ΩΛ=0 Ω tot= 1 0.5 1.0 ΩM 1.5 2.0 2.5 68.3 % 95.4% 99.7% Λ Ω 1 0 MLCS -1 0.0 Figure 5. The two parameter confidence contours are shown for the range of cosmological parameters ΩM and ΩΛ, given the current SNIa data from the High-Z collaboration. The solid and dotted contours are the constraints with and without the Fall 99 data, demonstrating the value of even a few points at z > 0.9. SNIa, H0, Λ, and t0 7 Figure 4 shows that a systematic effect that just grows as the redshift is not a good fit to the data, but the most telling way to separate a systematic effect that is proportional to redshift or time is to look at redshifts above 1. For a Λ-dominated cosmology, there is a transition, somewhere around z ∼ 1, from acceleration here and now to deceleration in the distant past when the matter density, which scales as (1 + z)3 would have been more important. The High-Z Team is working hard to test these ideas. Our Fall 1999 data, which are being slowly beaten into photometric perfection, emphasize z∼1 objects which will provide strong evidence to distinguish cosmology from systematics. Preliminary reductions are shown here. Our Fall 2000 program emphasized careful UBVRI photometry in the supernova rest frame to discern the effects of not-quite-gray dust. The final reduction of those data requires observations of the galaxy without the supernova, obtained a year after the discovery, and will be forthcoming when these templates are in hand (Jha 2001). 5. ΩM -- ΩΛ constraints and the Distribution of H0 t0 We can employ the tools of likelihood analysis to construct contours in the ΩM -- ΩΛ plane, as illustrated in Figure 5. Of special note is the effectiveness of just a handful of z > 0.9 supernovae in contracting the contours of this plot (as imagined by Goobar and Perlmutter 1995). A concise way to express the best constraint on ΩM and ΩΛ is ΩΛ − 1.6 ΩM = 0.4 ± 0.2. Interestingly, contours of constant age are very nearly parallel to the long axis of the error ellipse in Figure 5. This means that the competition between deceleration due to ΩM and acceleration due to ΩΛ is well captured by the measurement of luminosity distances, and the value of H0 t0 is well constrained even though the individual values of ΩM and ΩΛ are not. Marginalizing the probabilities of Figure 5 onto the H0 t0 axis yields a prob- ability distribution for H0 t0 as shown in Figure 6. We find H0 t0 = 1.00 ± 0.07. The absolute value of the cosmic age depends on H0, which, given the excellent agreement of the SBF and SNIa distances, inherits almost all its errors from the Cepheid zero point. 6. Summary Precise distance estimators measured over the range from z∼ 0 to z∼1 provide powerful constraints on the dimensionless product H0 t0 = 1.00 ± 0.07. Despite an ongoing struggle between deceleration and acceleration, the fortuitous result is that the formulation in elementary textbooks; t0 = 1/H0 turns out to be accurate. A more formal way to express the contraint from SNIa at z > 1 is that they imply ΩΛ − 1.6 ΩM = 0.4 ± 0.2. For our guess at the true value of H0, t0 = 14 ± 1.6 Gyr. A better value for the age of the Universe hinges on nearby matters like the distance to the LMC and metallicity dependence of Cepheids! References Aguirre, A. 1999 ApJ, 525, 583. SNIa, H0, Λ, and t0 8 Figure 6. The left panel shows the probability density distribution for H0 t0 using our data. The right panel shows the resulting distri- bution for t0, using the 2:1 normalization of the SNIa zero point to H0 = 73 and 65 suggested by the HST Cepheid calibrations. The un- certainty in H0 t0 (∼7%) is considerably smaller than the uncertainty H0 (∼ 12%) so the uncertainty in t0 comes mainly from the H0 and Cepheid zero point. Aguirre, A, and Haiman, Z. 2000 ApJ, 532, 28. Ajhar, E.A., Tonry, J.L., Blakeslee, J.P., Riess, A.G. & Schmidt, B.P. 2001, ApJ, in press. Bartlett, J.G. Blanchard, A. Silk, J. and Turner, M.S. 1995 Science 267, 980. Blakeslee, J.P., Ajhar, E.A., & Tonry, J.L. 1999, in Post-Hipparcos Cosmic Can- dles, eds. A. Heck & F. Caputo (Boston: Kluwer Academic Publishers), 181. Coil, A. et al. 2000, ApJ, 544, L111. Ferrarese, L. et al. 2000, ApJ, 529, 745. Filippenko, A. and Riess, A.G. (1998) Phys.Rept. 307, 31-44. Freedman, W. L., et al. 2001, ApJ, in press (astro-ph/0012376). Garnavich, P. M., et al. 1998, ApJ, 493, 53. Goobar, A. & Perlmutter, S. 1995 ApJ, 450, 14. Gibson, B. K. & Stetson, P.B. 2001 ApJ, 547, L103. Hamuy,M. et al. 1996 AJ, 112, 2398. Jha, S. et al. 1999 ApJS, 125, 73. Jha, S. et al. 2001 astro-ph/0101521. Leibundgut, B. et al. 1996 ApJ, 466, 21. Norgaard-Neilsen et al. 1989 Nature 339, 523. Parodi, B.R., Saha, A., Sandage, A., & Tammann, G.A. 2000, ApJ, 540, 634. Perlmutter et al. 1995, ApJ, 44, L41. SNIa, H0, Λ, and t0 9 Perlmutter et al. 1997, ApJ, 483, 565. Perlmutter et al. 1998, Nature 391, 54. Perlmutter et al. 1999, ApJ, 517, 565. Phillips, M. M. 1993, ApJ, 413, L105. Riess, A. G., Press, W. H., & Kirshner, R. P. 1996, ApJ, 473, 88. Riess et al. 1998 AJ, 116, 1009. Riess et al. 2000 ApJ, 536, 62. Riess et al. 2001 astro-ph/0104455. Saha, A. et al. 1997, ApJ, 486, 1. Schmidt et al. 1998 ApJ, 507, 46. Acknowledgments. Support for this work was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the As- sociation of Universities for Research in Astronomy, Inc. under NASA contract NAS5-26555.
0810.5120
1
0810
2008-10-28T20:00:03
Absolute dimensions of the F-type eclipsing binary star VZ Cephei
[ "astro-ph" ]
We present new V-band differential photometry and radial-velocity measurements of the unevolved 1.18-day period F+G-type double-lined eclipsing binary VZ Cep. We determine accurate values for the absolute masses, radii, and effective temperatures as follows: M(A) = 1.402 +/- 0.015 M(Sun), R(A) = 1.534 +/- 0.012 R(Sun), T(eff) = 6690 +/- 160 K for the primary, and M(B) = 1.1077 +/- 0.0083 M(Sun), R(B) = 1.042 +/- 0.039 R(Sun), T(eff) = 5720 +/- 120 K for the secondary. A comparison with current stellar evolution models suggests an age of 1.4 Gyr for a metallicity near solar. The temperature difference between the stars, which is much better determined than the absolute values, is found to be about 250 K larger than predicted by theory. If all of this discrepancy is attributed to the secondary (which would then be too cool compared to models), the effect would be consistent with similar differences found for other low-mass stars, generally believed to be associated with chromospheric activity. However, the radius of VZ Cep B (which unlike the primary, still has a thin convective envelope) appears normal, whereas in other stars affected by activity the radius is systematically larger than predicted. Thus, VZ Cep poses a challenge not only to standard theory but to our understanding of the discrepancies in other low-mass systems as well.
astro-ph
astro-ph
Accepted for publication in The Astronomical Journal Preprint typeset using LATEX style emulateapj v. 08/22/09 ABSOLUTE DIMENSIONS OF THE F-TYPE ECLIPSING BINARY STAR VZ CEPHEI Guillermo Torres1, and Claud H. Sandberg Lacy2 Draft version June 16, 2018 ABSTRACT We present new V -band differential photometry and radial-velocity measurements of the unevolved 1.18-day period F+G-type double-lined eclipsing binary VZ Cep. We determine accurate values for the absolute masses, radii, and effective temperatures as follows: MA = 1.402 ± 0.015 M⊙, RA = 1.534 ± 0.012 R⊙, Teff = 6690±160 K for the primary, and MB = 1.1077±0.0083 M⊙, RB = 1.042±0.039 R⊙, Teff = 5720 ± 120 K for the secondary. A comparison with current stellar evolution models suggests an age of 1.4 Gyr for a metallicity near solar. The temperature difference between the stars, which is much better determined than the absolute values, is found to be ∼250 K larger than predicted by theory. If all of this discrepancy is attributed to the secondary (which would then be too cool compared to models), the effect would be consistent with similar differences found for other low- mass stars, generally believed to be associated with chromospheric activity. However, the radius of VZ Cep B (which unlike the primary, still has a thin convective envelope) appears normal, whereas in other stars affected by activity the radius is systematically larger than predicted. Thus, VZ Cep poses a challenge not only to standard theory but to our understanding of the discrepancies in other low-mass systems as well. Subject headings: binaries: eclipsing -- stars: evolution -- stars: fundamental parameters -- stars: individual (VZ Cep) 1. INTRODUCTION VZ Cephei (also known as BD +70 1199 and GSC 04470-01334; α = 21h 50m 11.s14, δ = +71◦ 26′ 38.′′3, J2000.0; V = 9.72) was discovered photographically as a variable star by Gengler et al. (1928), who classified it to be of type "Is?", a rapid irregular variable. Cannon (1934) made the first spectral type assignment as G0. Its discovery as an eclipsing binary is due to Rossiger (1978), who determined a period of 1.18336 days and showed it to have unequal minima. The system was included by Lacy (1992, 2002a) in his photometric surveys of eclips- ing binary stars. Because of its relatively late spectral class, Popper (1996) had it as a target in his program to search for late-type (F -- K) eclipsing binary stars. He con- cluded on the basis of 4 spectrograms that both stellar components were likely hotter than G0. No determina- tions have been made of the physical properties of the stars, and the system has generally been neglected ex- cept for measurements of the times of eclipse made by a number of investigators since 1994. We began our investigation for the same reason Pop- per did: to test theoretical predictions of the properties of low-mass stars. We and other authors have previously found that in some of these binary systems the abso- lute properties are not well described by standard stellar evolution theory (see, e.g., Popper 1997; Clausen et al. 1999; Ribas 2006; Torres et al. 2006). We find below that VZ Cep also shows some anomalies compared to standard models, even though its components are both more massive than the Sun. 2. ECLIPSE EPHEMERIS 1 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, e-mail: [email protected] 2 Department of Physics, University of Arkansas, Fayetteville, AR 72701, e-mail: [email protected] Photometric times of minimum light of VZ Cep avail- able from the literature are collected in Table 1. Eclipse ephemerides determined by weighted least squares sepa- rately from the 15 primary minima and the 13 secondary minima gave the same period within the errors. A joint fit of all the measurements was then performed enforc- ing a common period but allowing the primary and sec- ondary epochs to be free parameters. Scale factors for the internal errors were determined by iterations sep- arately for the two types of measurements in order to achieve reduced χ2 values near unity. This solution re- sulted in a phase difference between the two epochs of ∆φ = 0.50033 ± 0.00022, not significantly different from 0.5. For our final ephemeris we imposed a circular orbit, and obtained Min I (HJD) = 2,452,277.324478(59) + 1.183363762(84) · E. The uncertainties of the fitted quantities in terms of the least significant digit are shown in parentheses. The final scale factors for the published internal errors were similar to those in the previous fit, and were 1.38 for the primary and 2.42 for the secondary. We adopt this ephemeris for the spectroscopic and photometric analyses below. 3. SPECTROSCOPIC OBSERVATIONS AND ORBIT VZ Cep was placed on the observing list at the Harvard-Smithsonian Center for Astrophysics (CfA) in 2003 January, and was observed until 2007 June with an echelle spectrograph on the 1.5-m Tillinghast reflector at the F. L. Whipple Observatory (Mount Hopkins, AZ). A single echelle order 45 A wide centered at 5188.5 A was recorded with an intensified Reticon photon-counting diode array, at a resolving power of ∆λ/λ ≈ 35,000. The strongest lines in this window are those of the Mg I b triplet. A total of 39 spectra were obtained with signal- to-noise ratios ranging from 19 to 47 per resolution ele- ment of 8.5 km s−1. 2 Torres & Lacy cross-correlation Radial velocities were measured with the two- technique TODCOR dimensional (Zucker & Mazeh 1994). Templates for the primary and secondary were selected from a large library of synthetic spectra based on model atmospheres by R. L. Kurucz3 (see also Nordstrom et al. 1994; Latham et al. 2002). These calculated spectra cover a wide range of effective temperatures (Teff ), rotational broadenings (v sin i when seen in projection), surface gravities (log g), and metal- licities. Solar metallicity was assumed throughout, along with initial values of log g = 4.5 for both components. The temperatures and rotational velocities for the templates were determined by running extensive grids of two-dimensional cross-correlations and seeking the best correlation value averaged over all exposures, as described in more detail by Torres et al. (2002). The secondary component in VZ Cep is some 5 times fainter than the primary, and we were unable to determine its temperature independently. We therefore adopted the results from other estimates described below, and chose the nearest value in our library, which is 5750 K. Due to the narrow wavelength range of our spectra the derived temperatures are strongly correlated with the assumed surface gravities. The secondary log g presented in § 5 is very close to the value we assumed, but the primary log g is intermediate between 4.0 and 4.5, so we repeated the calculations above using the lower value, and interpolated. The results for the primary are Teff = 6690 ± 150 K and v sin i = 57 ± 3 km s−1, and for the secondary we obtain v sin i = 50 ± 10 km s−1. Radial velocities were derived with template parameters near these values. The stability of the zero-point of the CfA velocity system was monitored by means of exposures of the dusk and dawn sky, and small run-to-run corrections were applied in the manner described by Latham (1992). Possible systematics in the radial velocities that may result from residual line blending in our narrow spectral window, or from shifts of the spectral lines in and out of this window as a function of orbital phase, were investi- gated by performing numerical simulations as described by Torres et al. (1997, 2000). Briefly, we generated ar- tificial composite spectra by adding together copies of the two templates with scale factors in accordance with the light ratio reported below, and with Doppler shifts for each star appropriate for each actual time of obser- vation, computed from a preliminary orbital solution. These simulated spectra were then processed with TOD- COR in the same manner as the real spectra, and the input and output velocities were compared. Experience has shown that the magnitude of these effects is diffi- cult to predict, and must be studied on a case-by-case basis. Corrections were determined for VZ Cep based on these simulations and were applied to the raw veloci- ties. The corrections for the primary star are small (un- der 1 km s−1), but for the secondary they are as large as 11 km s−1, and as expected they vary systematically with orbital phase or radial velocity (see Figure 1). Similarly large corrections have been found occasionally for other systems using the same instrumentation (e.g., AD Boo, GX Gem; Clausen et al. 2008; Lacy et al. 2008). The im- pact of these corrections is quite significant for VZ Cep: 3 Available at http://cfaku5.cfa.harvard.edu. Fig. 1. -- Corrections for systematics in the radial-velocity mea- surements for VZ Cep as a function of orbital phase and radial velocity (see text). Filled circles correspond to the primary and open circles to the secondary. the minimum masses increase by 4% for the primary and 1.9% for the secondary, and the mass ratio decreases by 2.1%. The final velocities in the heliocentric frame, in- cluding the corrections for systematics, are listed in Ta- ble 2 and have typical uncertainties of 1.3 km s−1 for the primary and 3.8 km s−1 for the fainter secondary. Preliminary single-lined orbital solutions using the pri- mary and secondary velocities separately indicated a zero-point difference between the two data sets (i.e., a difference in the systemic velocity γ), which is of- ten seen by many investigators in cases where there is a slight mismatch between the templates used for the cross-correlations and the spectra of the real stars (see, e.g., Popper 2000; Griffin et al. 2000). Numerous tests with other templates did not remove the offset.4 This most likely arises in our case because of stellar parame- ters (particularly the rotation) that fall in between the template parameters available in our library of synthetic spectra, which come in rather coarse steps of 10 km s−1 at the high rotation rates of VZ Cep. We therefore in- cluded this velocity offset as an additional free param- eter in the double-lined orbital fit, and we verified that when doing so the velocity semi-amplitudes (which de- termine the masses) are insensitive to the exact template parameters within reasonable limits, and are essentially identical to those resulting from separate single-lined so- lutions. Our final orbital fit is presented in Table 3. No indication of eccentricity was found, as expected for such a short period, so only a circular orbit was considered in the following. A graphical representation of the observa- 4 As a further test, solutions without applying the corrections for systematics described in the preceding paragraph gave an offset more than twice as large. VZ Cep 3 Fig. 2. -- Radial-velocity measurements for VZ Cep (including the corrections for systematics described in the text) along with our spectroscopic orbital solution. Filled circles correspond to the primary, and the dotted line represents the center-of-mass velocity. Error bars are smaller than the size of the points. The O − C residuals are shown on an expanded scale in the bottom panels, where typical error bars are indicated in the upper right corner. tions and our best fit, along with the residuals, is shown in Figure 2. The light ratio between the primary and secondary was determined from our spectra following Zucker & Mazeh (1994), accounting for the difference in line blocking be- tween the primary and the much cooler secondary. After corrections for systematics analogous to those described above, we obtained LB/LA = 0.19 ± 0.02 at the mean wavelength of our observations. A further adjustment to the visual band taking into account the temperature dif- ference between VZ Cep A and B was determined from synthetic spectra integrated over the V passband and the spectral window of our observations, and resulted in (LB/LA)V = 0.22 ± 0.02. 4. PHOTOMETRIC OBSERVATIONS AND ANALYSIS Differential photometry of VZ Cep was obtained at the URSA Observatory on the University of Arkansas campus at Fayetteville, AR. The URSA Observatory sits atop Kimpel Hall and consists of a Meade f/6.3, 10-inch Schmidt-Cassegrain telescope with a Santa Barbara In- struments Group ST8EN CCD camera inside a Technical Innovations RoboDome, all controlled by a Macintosh G4 computer in an adjacent control room inside the build- ing. The field of view is 20′ × 30′. Images of VZ Cep (V ≈ 9.7) and the two comparison stars GSC 04470- 01497 (V ≈ 9.9) and GSC 04470-01622 (V ≈ 11.2), both of which are within 6′ of the target, were taken with typ- ical integration times of 60 seconds through a Bessell V filter. With an overhead of about 30 seconds to down- load the images from the camera, the observing cadence was typically 90 seconds per image. A "virtual mea- Fig. 3. -- V -band photometric measurements for VZ Cep, along with our best constrained model described in the text. O−C resid- uals are shown at the bottom. suring engine" application written by Lacy was used to determine the brightness of the variable and comparison stars, to subtract off the sky brightness, and to correct for differences in airmass. A total of 5473 images were gathered between 2001 March 5 and 2003 September 7. Differential magnitudes were formed between the vari- able star and the magnitude corresponding to the sum of the fluxes of the two comparison stars. These are listed in Table 4, and shown graphically in Figure 3 along with our modeling described below. Expanded views of the primary and secondary eclipse are given in Figure 4 and Figure 5. The typical precision of these measurements is about 0.007 mag, which is comparable to that expected from photon statistics (∼0.006 mag). The comparison stars are not known to be variable. The mean magnitude difference between the two was constant with a standard deviation of 0.0095 mag over 67 nights, which is what would be expected for individual magnitudes with a typ- ical error of 0.007 mag. A Lomb-Scargle periodogram analysis of the individual differences was performed to search for periodic signals in either star, but none were detected. We have previously found (Lacy et al. 2008) that the URSA photometry is significantly improved by removal of small nightly zero-point variations. Thus 67 nightly corrections were made to the original magnitudes based on a preliminary photometric orbital fit. This procedure reduced the residual standard deviation by about 15%, a small but significant amount. Examination of these off- sets, which are typically smaller than 0.01 mag, revealed no detectable pattern as a function of orbital phase. Such a pattern might be expected, for instance, if there were perturbations in the light curve due to spots on either star (assuming synchronous rotation). We thus consider the nightly offsets to be instrumental in nature. The corrected photometry was fitted with the NDE model (Etzel 1981; Popper & Etzel 1981), with all ob- servations being assigned equal weight. In this model 4 Torres & Lacy Fig. 4. -- Enlarged view of Figure 3 showing the V -band pho- tometry for VZ Cep around the primary minimum. O−C residuals are shown at the bottom. Fig. 5. -- Enlarged view of Figure 3 showing the V -band photom- etry for VZ Cep around the secondary minimum. O −C residuals are shown at the bottom. the stars are represented as biaxial ellipsoids, and de- spite the relatively large radius of the primary of VZ Cep relative to the separation (see below), its ellipticity of 0.016, as defined by Etzel (1981), is still well below the maximum tolerance of 0.04 (Popper & Etzel 1981), and thus the model is expected to be adequate for this case. We return to this below. We used the JKTEBOP imple- mentation of Southworth et al. (2007) with a linear limb- darkening law, consistent with our findings (Lacy et al. 2005, 2008) that with the amount and precision of our data, non-linear limb-darkening laws do not improve the Fig. 6. -- Application of the constraint given by the spectroscopic brightness ratio to the light curve fits of VZ Cep. Grids of solu- tions for fixed values of k are shown for several key parameters, along with the corresponding rms residual of the fit (σ). The spec- troscopic value (LB/LA)V = 0.22 ± 0.02 is applied in the top left panel to determine k, and all other quantities are interpolated to the same value. accuracy of the fits significantly. The following quan- tities were allowed to vary in this unconstrained solu- tion: the central surface brightness JB of the secondary (cooler) star relative to the primary, the sum of relative radii rA + rB, the ratio of radii rB/rA, the orbital in- clination i, the limb-darkening coefficients uA and uB, a phase offset, and the magnitude at quadrature. The fol- lowing quantities were kept fixed: the orbital eccentric- ity e = 0, the mass ratio from the spectroscopic solution q ≡ MB/MA = 0.7900, and the gravity brightening expo- nents yA = 0.25 and yB = 0.36, set by the temperatures and surface gravities following Claret (1998). The uncer- tainties of the adjustable parameters were estimated with a Monte Carlo technique in which we generated 500 syn- thetic light curves, solved for the parameters, and calcu- lated the standard error of each parameter. This process yielded uncertainty estimates accurate to two significant digits, which is sufficient for our purposes. These "Un- constrained" results are given in Table 5. Tests allowing for non-zero eccentricity and third light gave statistically insignificant values for those quantities. The V -band light ratio (LB/LA)V from this fit is consistent with our spectroscopic value from § 3, but formally less precise. In similar systems with partial eclipses, the accuracy of the parameters (more than their precision) is sometimes compromised because of strong correlations among variables and the relatively flat bot- tom of the χ2 surface in the least-squares problem (see, e.g., Andersen 1991). In such systems it is often the case that more accurate results are obtained by applying the spectroscopic light ratio as an external constraint. We have done this here by first computing a grid of solutions VZ Cep 5 for a range of fixed values of k. We then interpolated in the smooth relation obtained between the light ratio and k to our value of (LB/LA)V = 0.22 ± 0.02, and derived k = 0.680±0.030. Interpolations of all other quantities to this value of k were then carried out. This is illustrated in Figure 6 for some of the key light-curve parameters. Note that σ, the rms residual of the fit, changes very little for k between 0.65 and 0.75, demonstrating that the radius ratio cannot be determined accurately from photometry alone without an external constraint. The results for the light curve parameters from this constrained fit are listed in Table 5, and are adopted for further use. The uncer- tainties have been propagated directly from the error in the spectroscopic light ratio (see Figure 6), and include also a contribution from the statistical uncertainties de- rived from a separate solution in which k was fixed at the best-fit value and all other parameters were left free. The fitted linear limb-darkening coefficients from this constrained solution tend to be somewhat smaller than predicted by theory. We find, for example, marginal dif- ferences of 0.10 (∼1.3σ) for both stars compared to the calculations by van Hamme (1993), and more significant differences of 0.19 (2.5σ) compared to the coefficients by Claret (2000). These differences are similar in magnitude (and in this particular case, of the same sign) as those reported, e.g., by Southworth (2008), and may be due to shortcomings in the theoretical model atmospheres al- though observational errors cannot be ruled out. For further comparisons between theory and observations the reader is referred to the recent work of Claret (2008). Adopting coefficients from the tables by Claret (2000) leads to values of the relative radii that are larger by 1.1% for the primary and 1.7% for the secondary (1.6σ and 0.5σ, respectively). As a test of the reliability of the geometric parameters, we carried out solutions with two other light-curve mod- eling programs that are more sophisticated than the one we have used. One is the WINK program (Wood 1972), which adopts a better approximation to the stellar shapes as triaxial ellipsoids, rather than the simpler biaxial ellip- soids in EBOP, and includes a more detailed treatment of reflection effects. The version we used has been modi- fied and extended as described by Vaz (1984, 1986), Vaz (1985), and Nordlund & Vaz (1990). The other program is the Wilson-Devinney code (WD; Wilson & Devinney 1971; Wilson 1979, 1990, 1993; van Hamme 2007) in its most recent (2007) release, which uses Roche geome- try. Light curve solutions with these two codes were performed for a fixed value of k = 0.680 (as closely as allowed by the different input quantities) to permit a direct comparison with our constrained JKTEBOP fit, and with the same limb-darkening law and coefficients as used earlier. The WINK fit delivered marginally smaller relative radii that differ from our JKTEBOP results by ∆rA = −0.0009 and ∆rB = −0.0005 (i.e., less than 0.4%), and an inclination angle that was only ∆i = +0.◦03 larger. The WD fit gave ∆rA = −0.0008, ∆rB = −0.0005, and ∆i = +0.◦22. These results are thus not significantly different from those of the simpler model we have used, as expected from the relatively small ellipticity of the stars mentioned earlier. The individual temperatures were determined from the central surface brightness parameter JB slightly ad- justed for limb darkening to correspond to the disk av- erage (see, e.g., Lacy et al. 1987), the absolute visual flux scale of Popper (1980), and an estimate of the mean system temperature used as the initial value for the primary. The latter was then improved by itera- tion until convergence. The mean system temperature is based on accurate Stromgren photometry for VZ Cep re- ported by Lacy (2002a). Interstellar reddening was esti- mated using the calibration of Perry & Johnston (1982) and the method of Crawford (1975), which resulted in E(b − y) = 0.032 ± 0.007 and an intrinsic color in- dex of (b − y)0 = 0.286 ± 0.007. The calibration by Holmberg et al. (2007) was then used to derive a mean system temperature of 6500 ± 150 K, assuming solar metallicity. The individual temperatures derived in this way are 6690 ± 160 K and 5720 ± 120 K for the pri- mary and secondary, respectively, which correspond to spectral types of approximately F3 and G4 (Gray 1992, p. 430). The primary Teff is identical to our spectro- scopic estimate in § 3. The temperature difference based the light curve is of course better determined than the absolute values: ∆Teff = 970 ± 35 K. Use of a different color/temperature calibration for inferring the mean sys- tem temperature, such as that by Alonso et al. (1996), yields results only ∼30 K hotter. 5. ABSOLUTE DIMENSIONS AND PHYSICAL PROPERTIES The spectroscopic and photometric solutions above lead to the masses and radii given in Table 6. Also in- cluded are the predicted projected rotational velocities, calculated under the assumption of synchronism with the orbital motion. The secondary value is consistent with our measured v sin i from § 3, but the expected value of 64.6 ± 0.5 km s−1 for the primary seems somewhat larger than our spectroscopic estimate of v sin i = 57±3 km s−1. At face value this would indicate sub-synchronous rota- tion of that component, which is unexpected in a short- period binary such as VZ Cep. Since the primary star dominates the light of the system, we investigated the possibility that there might be a photometric signal pro- duced, for instance, by rotational modulation from sur- face features on that component. For this we examined the residuals from our adopted light curve solution. A Lomb-Scargle power spectrum did not indicate any sig- nificant periodicities within a factor of two of the orbital frequency, although the primary star is likely to be too hot for spots to be important (see § 7). There are no measurements of the chemical abundance of VZ Cep. Our own spectroscopy is inadequate for this, and the combined-light Stromgren indices along with the calibration by Holmberg et al. (2007) indicate [Fe/H] = +0.06 ± 0.09, in which the uncertainties include pho- tometric errors as well as the scatter of the calibration. The Hipparcos catalog (Perryman et al. 1997) has no en- try for VZ Cep and no trigonometric parallax is available. From its radiative properties as measured here we find the distance to the system to be 215 ± 10 pc (similar distances are obtained separately for each component, indicating the consistency of the measured properties). Discrepancies described in the next section between our effective temperature estimates and the Teff val- ues predicted by stellar evolution models prompted us to attempt a deconvolution of the combined-light pho- tometry of VZ Cep, as a check on both the color ex- cess and the temperature difference between the com- 6 Torres & Lacy ponents. We used tables of standard Stromgren indices by Crawford (1975) and Olsen (1984), and synthesized binary stars for a range of primary indices and a fixed V -band light ratio given by our spectroscopic estimate of (LB/LA)V = 0.22 ± 0.02. We explored a wide range of E(b−y) values. At each reddening we determined the intrinsic indices for the primary and secondary that pro- vide the best match to the system values of b−y, m1, c1, and β as measured by Lacy (2002a), in the χ2 sense. We found the best agreement for E(b−y) = 0.032, in excel- lent accord with our earlier determination based on the combined light. The measured b−y, c1, and β indices are reproduced to well within their uncertainties, and m1 is within 1.8σ. Making use of the same color/temperature calibration by Holmberg et al. (2007) invoked earlier, the intrinsic indices for each star from this photometric de- convolution yield temperatures of 6670 K and 5690 K, once again in very good agreement with the light curve results. The temperature difference from this exercise is ∆Teff = 975 ± 40 K. 6. COMPARISON WITH STELLAR EVOLUTION THEORY The absolute masses of VZ Cep have formal relative errors of 1% or better. The radius of the primary is similarly well determined, while that of the faint sec- ondary is good to about 3.7%. These values along with the temperatures are compared here with stellar evolu- tion models from the Yonsei-Yale series (Yi et al. 2001; Demarque et al. 2004). In Figure 7 the measurements are shown in the R vs. Teff plane against evolutionary tracks computed for the measured masses and for solar metallicity (Z = 0.01812 in these models, indicated with solid lines). The shaded areas represent the uncertainty in the location of each track due to the measurement errors in the masses MA and MB. While the primary track is in good agreement with our temperature deter- mination for that star, the secondary track is too hot. Adjustment of the chemical composition of the models to Z = 0.0280 (corresponding to [Fe/H] = +0.21) provides the fit shown with the dotted lines. This fit is marginally consistent with our temperature error bars in the figure, but the agreement is misleading since the temperature difference is much better determined than the error bars appear to indicate. The best-fit age for this metallicity is 1.6 Gyr, and the corresponding isochrone is indicated with a dashed curve. Figure 8 shows the measurements in the mass-radius diagram against the same set of models. The dashed line represents the same isochrone shown before, and the solid line is an isochrone for solar metallicity that provides the best fit, in this case for a slightly younger age of 1.4 Gyr. Both are seen to represent the measurements equally well. These comparisons suggest that the models correctly predict the radii of the stars at the measured masses, but that the temperature of the secondary is underestimated by a significant amount. Tests with a different series of models by Pietrinferni et al. (2004) gave similar results. Fig. 7. -- Stellar evolution models from the Yonsei-Yale series (Yi et al. 2001; Demarque et al. 2004) compared against the mea- surements for VZ Cep. Solid lines show evolutionary tracks for the measured masses and for solar metallicity (Z = Z⊙), with the uncertainty in the location of the tracks represented by the shaded regions. Dotted lines correspond to mass tracks at a somewhat higher metallicity of Z = 0.0280 that seems to fit the observations better (see text). An isochrone for this metallicity and an age of 1.6 Gyr is shown with a dashed curve. Fig. 8. -- Isochrones from the Yonsei-Yale series (Yi et al. 2001; Demarque et al. 2004) compared with the measurements for VZ Cep in the mass-radius plane. The dashed line is the same isochrone shown in Figure 7 (Z = 0.0280), and a solar-metallicity isochrone is represented by the solid curve, for a slightly younger age that fits the observations best. 7. DISCUSSION AND CONCLUSIONS VZ Cep stands out among the F stars as one of the eclipsing binaries with the largest difference in mass be- tween the components (q = 0.7900), which provides ex- tra leverage for testing stellar evolution models. There are no less than five other systems with well deter- mined properties (BW Aqr B, V1143 Cyg A, BP Vul B, V442 Cyg B, and AD Boo A; Andersen 1991; Lacy et al. 2003; Clausen et al. 2008) that have at least one compo- nent nearly identical in mass to the primary of VZ Cep VZ Cep 7 (i.e., within 1%). However, these stars are all in very different evolutionary states so that their radii span a range of 33% and their effective temperatures differ by up to 360 K. They are therefore of little help in under- standing the discrepancies with theory noted above for VZ Cep. Only one other well measured binary has one component with a mass similar to that of the secondary of VZ Cep, but that star (EK Cep B) is considered to be in the pre-main sequence stage (Popper 1987). Figure 7 highlights the main disagreement between the models and the measurements for VZ Cep, which is that the temperature difference predicted from theory for the measured masses and surface gravities is much smaller than all of our estimates. Solar metallicity models, which appear to fit the properties of the primary well, indicate ∆Teff = 710 K, and this is reduced further to 660 K when considering the higher metallicity of Z = 0.0280. The uncertainty in these determinations is difficult to quantify, but a useful measure may be obtained by prop- agating the uncertainty in the measured masses, which results in ±50 K. Uncertainties from physical inputs to the models are unlikely to add much to this due to the dif- ferential nature of the comparison. In this paper we have made three empirical determinations of ∆Teff, as follows: 1) ∆Teff = 970 ± 35 K, based on the JB value from our light-curve analysis along with the visual flux scale of Popper (1980) and our spectroscopic brightness ratio (used as an external constraint); 2) ∆Teff = 975 ± 40 K, from photometric deconvolution based on the measured Stromgren indices and the spectroscopic brightness ratio (§ 5), along with the color/temperature calibrations of Holmberg et al. (2007); 3) ∆Teff = 940 K, directly from a primary temperature estimate based on spectroscopy (§ 3) and an assumed temperature for the secondary sim- ilar to estimates for that star from the other two meth- ods. While these three determinations are not completely independent, their consistency despite the widely differ- ent ingredients reinforces our conclusion that the model ∆Teff is at least ∼250 K too small. Stellar evolution models have been shown previously to overestimate the effective temperatures of low-mass stars in eclipsing binaries by up to ∼200 K (e.g., Torres & Ribas 2002; Ribas 2003). The study of V1061 Cyg by Torres et al. (2006) suggested that the problem is not confined to M dwarfs, but extends up to masses almost as large as that of the Sun (0.93 M⊙ in the case of V1061 Cyg B). At the same time, the radii of these stars appear too large compared to theory, and both discrepancies are generally attributed to the effects of stellar activity in these short-period, tidally synchro- nized and rapidly rotating systems. There is little doubt that the VZ Cep system is ac- tive, judging by its strong X-ray emission as recorded by ROSAT (Voges et al. 1999). We estimate its X-ray luminosity to be log LX = 30.61 ± 0.12 (where LX is in cgs units), and log LX/Lbol = −3.70 ± 0.13.5 The mass of VZ Cep B is slightly larger than that of the Sun, but it still has a thin convective envelope representing about 1.3% of the total mass, which suggests that star may in fact be responsible for most of the X-ray emission given 5 By comparison, log LX for the Sun ranges between 26.4 and 27.7 during the activity cycle (Peres et al. 2000), and log LX/Lbol ranges between −7.2 and −5.9. that the primary has no convective envelope. Another in- dication is given by the Rossby numbers of the stars (ra- tio R0 between the convective turnover time and the rota- tional period). For the primary we estimate log R0 > 2.1, which according to Hall (1994, Fig. 6) clearly places it among the inactive stars. The secondary, on the other hand, has log R0 = −1.3. Stars in this regime tend to be very active and have photometric variability due to spots with amplitudes as large as ∼0.4 mag. Detection of this expected variability is difficult in VZ Cep because of the faintness of the secondary. Nevertheless, we exam- ined the nightly residuals from our adopted solution near the primary minimum, where the contrast is more favor- able, and we see only occasional systematic deviations on one or two nights. However, similar deviations are seen at the secondary eclipse, and also outside of eclipse, which leads us to believe these are residual instrumental errors rather than real changes in the light level caused by spottedness on the secondary (see § 4). If we consider the properties of the primary of VZ Cep to be relatively well described by theory for a metal- licity near solar, then the secondary shows a temper- ature difference compared to models in the same di- rection as mentioned above for other active stars (i.e., lower than predicted). However, we see no indication that its radius is larger than predicted (Figure 8), which we would have expected not only from the evidence dis- played by other systems but also from recent theoreti- cal studies of the effects of chromospheric activity (e.g., Mullan & MacDonald 2001; Chabrier et al. 2007). As described in previous sections, we have carried out a variety of tests to explore the possibility of systematic er- rors in our mass, radius, or temperature determinations, including a careful examination of biases in our velocity measurements, and sanity checks of our light-curve anal- ysis with results from more sophisticated modeling pro- grams. We were unable to demonstrate any significant errors that would explain the discrepancies in the preced- ing paragraph. For example, matching the model ∆Teff with the ∆Teff measured from the light curve would re- quire a change in the mass ratio to q ≈ 0.71, much lower than allowed by the spectroscopy, regardless of the choice of cross-correlation templates (see § 3). Conversely, de- riving a smaller temperature difference from the light curve to match the models would require an increase in JB to values that are unrealistic and would bring strong disagreement with the light ratio from spectroscopy. Ad- ditionally, this would leave the other two empirical esti- mates of ∆Teff unchanged, and a discrepancy would re- main. As indicated earlier, we see no evidence for third light at a level that would make much difference. The ad- justments required in each of the quantities mentioned above, and others we experimented with, are so large compared to the uncertainties that a combination of ef- fects is unlikely to resolve the issue either. At the suggestion of the referee, we show in Fig- ure 9 a comparison with an alternate set of models by Baraffe et al. (1998), which allows us to explore the ef- fect of differences in the mixing length parameter αML. Previous studies of the radius and temperature discrep- ancies for active low-mass stars have indicated that a value of αML lower than appropriate for the Sun, repre- senting a reduced overall convective efficiency, provides a much better match to the observations of these objects. 8 Torres & Lacy predicted, in agreement with our earlier conclusion that this star appears normal in size (compared to standard models). The 13% difference between our measured v sin i for the primary and the predicted synchronous velocity is somewhat puzzling, and is significant at the 2.5-σ confi- dence level. We do not believe errors in the spectroscopic measurements are to blame since all 39 of our individual spectra consistently give values smaller than predicted. A reduction in the predicted value could be accomplished with an increase in k, but it would have to be much larger than allowed by our photometric solutions, and would once again bring disagreement between the photometric and spectroscopic light ratios. At the moment we are unable to offer an explanation for the differences noted above, and based on the tests just described we are inclined to believe that the mea- surements are accurate and that the system is affected in some way that the models do not account for, most likely having to do with chromospheric activity. It would also appear that our understanding of the effects of chro- mospheric activity (reduced convective efficiency, spot coverage) on the global properties of stars is still incom- plete, since we see here only the effect on the tempera- ture predicted by recent models that account for these phenomena (Chabrier et al. 2007), but not the effect on the radius. VZ Cep thus presents a challenge to theory. Further progress in understanding these differences may be made by obtaining complete light curves in multiple passbands, which would give a better handle on the tem- perature issue. Higher signal-to-noise ratio spectroscopy would also help in refining the v sin i measurements for the primary and secondary, in constraining the abun- dance, and perhaps also in providing a more direct de- termination of the effective temperatures and revealing whether Ca II emission is present. The spectroscopic observations of VZ Cep used in this paper were obtained with the expert assistance of P. Berlind, M. Calkins, D. W. Latham, and R. P. Stefanik. R. J. Davis is thanked for maintaining the CfA echelle database. We are grateful to the referee, J. V. Clausen, for a prompt, detailed, and very helpful report. GT ac- knowledges partial support for this work from NSF grant AST-0708229. CHSL would like to thank University of Arkansas graduate student Kathryn D. Hicks for a pre- liminary analysis of the photometry and radial velocities of VZ Cep. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France, and of NASA's Astrophysics Data System Abstract Service. Fig. 9. -- Radius and effective temperature of VZ Cep compared against evolutionary tracks for solar metallicity by Baraffe et al. (1998), for the exact masses we measure (solid lines). A single track is shown for the primary star, for a mixing length parameter αML = 1.9 appropriate for the Sun. The dashed line represents the same solar-metallicity Yonsei-Yale track shown in Figure 7. Three Baraffe et al. (1998) models are shown for the secondary, for different values of αML, as labeled. For reference, the triangles on the solid curves correspond to an age of 1.4 Gyr, which was found in Figure 8 to provide a good match in the mass-radius plane using solar-metallicity models from the Yonsei-Yale series. Consequently, we show a solar-metallicity evolutionary track for a solar-like mixing length parameter for the ra- diative primary (αML = 1.9 in these models), and tracks for three values of the mixing length parameter for the secondary star (αML = 1.9, 1.5, and 1.0), which, as men- tioned earlier, we believe to be the more active member of the system. For reference, triangles on each track mark the age of 1.4 Gyr, which we found in Figure 7 to provide the best match for Z = Z⊙ using the Yonsei-Yale mod- els. The Baraffe et al. (1998) model for the primary is seen to be similar to the corresponding solar-metallicity Yonsei-Yale model (dashed line, reproduced from Fig- ure 7). A reduction of the mixing length parameter for the secondary star leads to the expected systematic de- crease in effective temperature, and an increase in ra- dius. A secondary model with αML between 1.0 and 1.5, when paired with the standard αML = 1.9 model for the primary, would appear to give approximately the cor- rect temperature difference for the system. However, the measured radius of VZ Cep B is somewhat smaller than REFERENCES Agerer, F., & Huebscher, J. 1995, IBVS, No. 4222 Alonso, A., Arribas, S., & Mart´ınez-Roger, C. 1996, A&A, 313, 873 Andersen, J. 1991, A&A Rev., 3, 91 Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 1998, A&A, 337, 403 Canon, A. J. 1934, Bull. Harvard Obs., 897, 12 Chabrier, G., Gallardo, J., & Baraffe, I. 2007, A&A, 472, L17 Claret, A. 1998, A&AS, 131, 395 Claret, A. 1998, A&A, 363, 1081 Claret, A. 2008, A&A, 482, 259 Clausen, J. V., Helt, B. E., & Olsen, E. H. 1999, in Theory and Tests of Convection in Stellar Structure, ASP Conf. Ser. 173, ed. A. Gim´enez, E. F. Guinan, & B. Montesinos (San Francisco: ASP), 321 Clausen, J. V., Torres, G., Bruntt, H., Andersen, J., Nordstrom, B., Stefanik R. P., Latham, D. W., & Southworth, J. 2008, A&A, 487, 1095 Crawford, D. L. 1975, AJ, 80, 955 Demarque, P., Woo, J.-H., Kim, Y.-C., & Yi, S. K. 2004, ApJS, 155, 667 Diethlem, R. 2006, IBVS, No. 5713 VZ Cep 9 Etzel, P. B. 1981, Photometric and Spectroscopic Binary Systems (Dordrecht: Reidel), 65 Gengler, G. T., Blasko, S., & Schneller, H. 1928, AN, 233, 39 Gray, D. F. 1992, The Observation and Analysis of Stellar Photospheres, 2nd Edition (Cambridge: Cambridge Univ. Press) Griffin, R. E. M., David, M., & Verschueren, W. 2000, A&AS, 147, 299 Hall, D. S. 1994, Mem. R. Astr. Soc., 65, 73 Holmberg, J., Nordstrom, B., & Andersen, J. 2007, A&A, 475, 519 Kim, C.-H., Lee, C.-U., Yoon, Y.-N., Park, S.-S., Kim, D.-H., Cha, S.-M., & Won, Y.-H. 2006, IBVS, No. 5694 Lacy, C. H., Freuh, M. L., & Turner, A. E. 1987, AJ, 94, 1035 Lacy, C. H. 1992, AJ, 104, 801 Lacy, C. H. S. 2002a, AJ, 124, 1162 Lacy, C. H. S. 2002b, IBVS, No. 5357 Lacy, C. H. S., Torres, G., Claret, A., & Sabby, J. A. 2003, AJ, 126, 1905 Lacy, C. H. S., Straughn, A., & Denger F. 2002, IBVS, No. 5251 Lacy, C. H. S., Torres, G., & Claret, A. 2008, AJ, 135, 1757 Lacy, C. H. S., Torres, G., Claret, A., & Vaz, L. P. R. 2005, AJ, 130, 2838 Latham, D. W. 1992, in IAU Coll. 135, Complementary Approaches to Double and Multiple Star Research, ASP Conf. Ser. 32, eds. H. A. McAlister & W. I. Hartkopf (San Francisco: ASP), 110 Latham, D. W., Stefanik, R. P., Torres, G., Davis, R. J., Mazeh, T., Carney, B. W., Laird, J. B., & Morse, J. A. 2002, AJ, 124, 1144 Mullan, D. J., & MacDonald, J. 2001, ApJ, 559, 353 Nelson, R. H. 2001, IBVS, No. 5040 Nelson, R. H. 2007, IBVS, No. 5760 Nordlund, A, & Vaz, L. P. R. 1990, A&A, 228, 231 Nordstrom, B., Latham, D. W., Morse, J. A., Milone, A. A. E., Kurucz, R. L., Andersen, J., & Stefanik, R. P. 1994, A&A, 287, 338 Olsen, E. H. 1984, A&AS, 57, 443 Peres, G., Orlando, S., Reale, F., Rosner, R., & Hudson, H. 2000, ApJ, 528, 537 Perry, C. L., & Johnston, L. 1982, ApJS, 50, 451 Perryman, M. A. C., et al. 1997, The Hipparcos and Tycho Catalogues (ESA SP-1200; Noordwjik: ESA) Pietrinferni, A., Cassisi, S., Salaris, M., & Castelli, F. 2004, ApJ, 612, 168 Popper, D. M. 1980, ARA&A, 18, 115 Popper, D. M. 1987, ApJ, 313, L81 Popper, D. M. 1996, ApJS, 106, 133 Popper, D. M. 1997, AJ, 114, 1195 Popper, D. M. 2000, AJ, 119, 2391 Popper, D. M., & Etzel, P. B. 1981, AJ, 86, 102 Ribas, I. 2006, Ap&SS, 304, 89 Ribas, I. 2003, A&A, 398, 239 Rossiger, S. 1978, IBVS, No. 1474 Sarounova, L., & Wolf, M. 2005, IBVS, No. 5594 Southworth, J. 2008, MNRAS, 386, 1644 Southworth, J., Bruntt, H., & Buzasi, D. L. 2007, A&A, 467, 1215 Torres, G., Andersen, J., Nordstrom, B., & Latham, D. W. 2000, AJ, 119, 1942 Torres, G., Neuhauser, R., & Guenther, E. W. 2002, AJ, 123, 1701 Torres, G., Stefanik, R. P., Andersen, J., Nordstrom, B., Latham, D. W., & Clausen, J. V. 1997, AJ, 114, 2764 Torres, G., Lacy, C. H. S., Marschall, L. A., Sheets, H. A., & Mader, J. A. 2006, ApJ, 640, 1018 Torres, G., & Ribas, I. 2002, ApJ, 567, 1140 van Hamme, W. 1993, AJ, 106, 2096 van Hamme, W., & Wilson, R. E. 2007, ApJ, 661, 1129 Vaz, L. P. R. 1984, Ph.D. Thesis, Copenhagen University (unpublished) Vaz, L. P. R. 1986, Rev. Mexicana Astron. Astrofis., 12, 177 Vaz, L. P. R., & Nordlund, A 1985, A&A, 147, 281 Voges, W. et al. 1999, A&A, 349, 389 Wilson, R. E., & Devinney, E. J. 1971, ApJ, 166, 605 Wilson, R. E. 1979, ApJ, 234, 1054 Wilson, R. E. 1990, ApJ, 356, 613 Wilson, R. E. 1993, in New Fronteers in Binary Star Research, ASP Conf. Ser. 38, ed. K.-C. Leung & I.-S. Nha (San Francisco: ASP), 91 Wood, D. B. 1972, A Computer Program for Modeling Non-Spherical Eclipsing Binary Star Systems, Goddard Space Flight Center, Greenbelt, Maryland Yi, S. K., Demarque, P., Kim, Y.-C., Lee, Y.-W., Ree, C. H., Lejeune, T., & Barnes, S. 2001, ApJS, 136, 417 Zucker, S., & Mazeh, T. 1994, ApJ, 420, 806 10 Torres & Lacy TABLE 1 Published measurements of the times of eclipse for VZ Cep. HJD (2,400,000+) Typea Uncertainty (days) (O −C) (days) Source 49567.42210 . . . 51608.72370 . . . 52038.87680 . . . 52044.79410 . . . 52051.89410 . . . 52054.85215 . . . 52073.78570 . . . 52076.74440 . . . 52079.70300 . . . 52080.88604 . . . 52093.90270 . . . 52108.69630 . . . 52111.65270 . . . 52112.83709 . . . 52114.61500 . . . 52154.84470 . . . 52159.58000 . . . 52166.67970 . . . 52179.69710 . . . 52233.54070 . . . 52277.32429 . . . 52463.70530 . . . 52464.88810 . . . 52482.63870 . . . 52518.73064 . . . 53366.01950 . . . 53658.30900 . . . 54009.76910 . . . 1 1 2 2 2 1 1 2 1 1 1 2 1 1 2 2 2 2 2 1 1 2 2 2 1 1 1 1 0.0003 0.0003 0.0005 0.0005 0.0005 0.00011 0.0002 0.0003 0.0003 0.00019 0.0005 0.0003 0.0006 0.00014 0.0006 0.0004 0.0010 0.0003 0.0003 0.0004 0.00007 0.0003 0.0003 0.0005 0.00019 0.0002 0.0006 0.0001 +0.00067 −0.00023 −0.00012 +0.00036 +0.00018 +0.00008 −0.00019 −0.00016 +0.00029 −0.00003 −0.00038 +0.00092 −0.00083 +0.00019 +0.00280 −0.00187 −0.00002 −0.00051 −0.00011 +0.00070 −0.00017 +0.00079 +0.00022 +0.00037 −0.00003 +0.00037 −0.00098 +0.00008 1 2 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 4 5 5 5 5 6 7 8 Note. -- References: 1. Agerer & Huebscher (1995); 2. Nelson (2001); 3. Lacy et al. (2002); 4. Sarounova & Wolf (2005); 5. Lacy (2002b); 6. Kim et al. (2006); 7. Diethlem (2006); 8. Nelson (2007). a Type: 1 = primary eclipse; 2 = secondary eclipse. VZ Cep TABLE 2 11 Radial velocity measurements of VZ Cep. HJD (2,400,000+) RVA (km s−1) RVB (km s−1) (O −C)A (O −C)B (km s−1) (km s−1) Orbital phase 52661.5709 . . . . 52808.9615 . . . . 52828.9269 . . . . 52834.9798 . . . . 52885.8243 . . . . 52894.8309 . . . . 52924.7873 . . . . 52951.7261 . . . . 52958.6957 . . . . 53013.5806 . . . . 53182.9448 . . . . 53185.9637 . . . . 53186.9164 . . . . 53189.9839 . . . . 53191.8776 . . . . 53192.9211 . . . . 53215.9246 . . . . 53218.8991 . . . . 53271.6999 . . . . 53272.7658 . . . . 53274.6998 . . . . 53275.7597 . . . . 53281.7191 . . . . 53282.8175 . . . . 53301.7530 . . . . 53333.7365 . . . . 53335.6589 . . . . 53336.6417 . . . . 53339.6764 . . . . 53630.7663 . . . . 53636.7401 . . . . 53690.6534 . . . . 53691.7072 . . . . 54042.6132 . . . . 54043.7068 . . . . 54074.5790 . . . . 54103.5713 . . . . 54279.9427 . . . . 54282.8950 . . . . +106.35 −149.88 −127.46 +139.25 −96.88 +102.39 −127.64 +147.21 −126.22 +134.87 +96.39 −136.82 −100.02 +112.30 +57.66 −102.06 +105.49 −155.58 −114.11 +129.47 −125.76 +138.43 +88.37 −131.20 +88.42 −133.24 −129.03 +137.70 +89.70 −128.04 +106.23 −154.75 −112.11 +119.86 +94.48 −137.57 −122.28 +137.26 −122.11 +134.75 +94.19 −144.23 +104.75 −162.66 +107.50 −158.17 +100.85 −144.52 +100.90 −147.17 +106.56 −150.44 −109.05 +117.57 −116.21 +124.74 +108.41 −157.55 +106.36 −156.85 +108.74 −160.82 −112.94 +127.71 −127.60 +134.58 +111.73 −160.65 +98.06 −145.42 +106.66 −154.30 −127.62 +139.12 −121.65 +138.43 +105.11 −148.54 +1.69 +1.12 +0.15 +1.10 −0.81 −0.31 +0.34 −0.97 +0.59 +0.78 −1.35 −0.31 +2.37 −0.41 0.00 −1.14 −0.73 −2.02 +1.64 −0.01 +0.17 −2.93 −1.13 +0.05 −0.28 −0.27 −0.46 −0.51 −0.30 −1.07 +0.93 +0.49 −0.62 +2.80 +0.68 −1.60 +0.44 −1.19 +3.09 +2.71 −3.40 −0.32 +4.36 −3.76 +5.69 +5.38 −7.73 −2.69 +4.16 +1.06 +1.16 −4.21 −4.99 +5.61 +1.27 −1.02 +4.69 +0.51 +0.31 −5.11 −6.25 −0.56 +3.18 +1.01 +4.89 +0.23 −1.60 +0.16 −0.75 −4.24 +4.24 −6.04 −2.66 −2.05 +2.85 −2.87 +6.07 +0.71 0.707 0.259 0.131 0.246 0.212 0.823 0.138 0.902 0.792 0.172 0.293 0.845 0.650 0.242 0.842 0.724 0.163 0.677 0.296 0.197 0.831 0.727 0.763 0.691 0.692 0.720 0.344 0.175 0.739 0.724 0.773 0.332 0.222 0.755 0.679 0.768 0.268 0.310 0.805 Note. -- These velocities include corrections for systematics (see text). 12 Torres & Lacy Spectroscopic orbital solution for VZ Cep. TABLE 3 Parameter Value Adjusted quantities P (days)a . . . . . . . . . . . . . . . . . 1.18336377 TI (HJD−2,400,000)a . . . . . 52,277.32446 KA (km s−1) . . . . . . . . . . . . . 118.88 ± 0.22 KB (km s−1). . . . . . . . . . . . . . 150.48 ± 0.67 γ (km s−1) . . . . . . . . . . . . . . . . −9.90 ± 0.21 ∆RV (km s−1)b . . . . . . . . . . −2.31 ± 0.65 Derived quantities MA sin3 i (M⊙) . . . . . . . . . . . MB sin3 i (M⊙) . . . . . . . . . . . q ≡ MB/MA . . . . . . . . . . . . . . aA sin i (106 km) . . . . . . . . . . aB sin i (106 km) . . . . . . . . . . a sin i (R⊙) . . . . . . . . . . . . . . . 1.339 ± 0.013 1.0577 ± 0.0064 0.7900 ± 0.0038 1.9345 ± 0.0036 2.4487 ± 0.0109 6.298 ± 0.016 Other quantities pertaining to the fit Nobs . . . . . . . . . . . . . . . . . . . . . . Time span (days) . . . . . . . . . σA (km s−1) . . . . . . . . . . . . . . σB (km s−1) . . . . . . . . . . . . . . 39 1621.3 1.27 3.82 a Ephemeris adopted from § 2. b Velocity offset in the sense hprimary−secondaryi (see text). Differential V -band magnitudes of TABLE 4 VZ Cep. HJD−2,440,000 ∆V Orbital phase 51973.95823 51973.95926 51973.96028 51973.96130 51973.96233 0.129 0.115 0.123 0.126 0.126 0.64077 0.64164 0.64250 0.64336 0.64423 Note. -- Table 4 is available in its entirety in the electronic edition of the Astronomical Journal. A portion is shown here for guidance regarding its form and contents. Photometric orbital solutions for VZ Cep. TABLE 5 Parameter Unconstrained fit Constrained fit JB . . . . . . . . . . . . k ≡ rB/rA . . . . . rA + rB . . . . . . . rA . . . . . . . . . . . . . rB . . . . . . . . . . . . . uA . . . . . . . . . . . . uB . . . . . . . . . . . . i (deg) . . . . . . . . LA(V )a . . . . . . . (LB/LA)V . . . . σV (mmag). . . . Nobs . . . . . . . . . . 0.495 ± 0.010 0.717 ± 0.041 0.4077 ± 0.0043 0.2374 ± 0.0033 0.1703 ± 0.0073 0.499 ± 0.075 0.581 ± 0.059 79.47 ± 0.44 0.802 ± 0.022 0.246 ± 0.033 7.4400 5473 0.4920 ± 0.0013 0.680 ± 0.030 0.4028 ± 0.0077 0.2398 ± 0.0017 0.1630 ± 0.0061 0.420 ± 0.076 0.500 ± 0.076 79.97 ± 0.45 0.820 ± 0.014 0.220 ± 0.020b 7.4365 5473 a Fractional luminosity of the primary. b Adopted as a constraint from spectroscopy (see text). 13 VZ Cep TABLE 6 Physical properties of VZ Cep. Parameter Primary Secondary Mass (M⊙) . . . . . . . . Radius (R⊙) . . . . . . . log g (cgs) . . . . . . . . . . Temperature (K) . . . log L (L⊙) . . . . . . . . . v sin i (km s−1)a . . . . vsync sin i (km s−1)b a (R⊙). . . . . . . . . . . . . Distance (pc) . . . . . . Mbol (mag) . . . . . . . . MV (mag) . . . . . . . . . 1.402 ± 0.015 1.534 ± 0.012 4.2130 ± 0.0080 6670 ± 160 0.634 ± 0.041 57 ± 3 64.6 ± 0.5 1.1077 ± 0.0083 1.042 ± 0.039 4.446 ± 0.033 5720 ± 120 0.026 ± 0.050 50 ± 10 43.9 ± 1.6 6.396 ± 0.019 215 ± 10 3.18 ± 0.10 3.16 ± 0.11 4.68 ± 0.12 4.77 ± 0.12 a Value measured spectroscopically. b Value predicted assuming synchronous rotation.
astro-ph/0512479
1
0512
2005-12-19T16:55:03
Cosmogenic Neutrinos from Cosmic Ray Interactions with Extragalactic Infrared Photons
[ "astro-ph" ]
We discuss the production of cosmogenic neutrinos on extragalactic infrared photons in a model of its cosmological evolution. The relative importance of these infrared photons as a target for proton interactions is significant, especially in the case of steep injection spectra of the ultrahigh energy cosmic rays. For an E$^{-2.5}$ cosmic ray injection spectrum, for example, the event rate of neutrinos of energy above 1 PeV is more than doubled.
astro-ph
astro-ph
Cosmogenic Neutrinos from Cosmic Ray Interactions with Extragalactic Infrared Photons Bartol Research Institute, University of Delaware, Newark, DE 19716, USA Todor Stanev∗ and Daniel De Marco† F.W. Stecker‡ NASA Goddard Space Flight Center (Dated: October 2, 2018) We discuss the production of cosmogenic neutrinos on extragalactic infrared photons in a model of its cosmological evolution. The relative importance of these infrared photons as a target for proton interactions is significant, especially in the case of steep injection spectra of the ultrahigh energy cosmic rays. For an E−2.5 cosmic ray injection spectrum, for example, the event rate of neutrinos of energy above 1 PeV is more than doubled. PACS numbers: 98.70.Sa,98.70.Lt,13.85.Tp,98.80.Es I. INTRODUCTION The assumption that the ultra high energy cosmic rays (UHECR) are nuclei (presumed here to be protons) accel- erated in powerful extragalactic sources provides a nat- ural connection between these particles and ultra high energy neutrinos. This was first realized by Berezinsky and Zatsepin [1] soon after the introduction of the GZK effect [2]. The GZK effect is the modification of the UHE proton spectrum from energy losses by photoproduction interactions with the 2.7K microwave background radi- ation (MBR). In the case of isotropic and homogeneous distribution of UHE cosmic ray sources, the GZK effect leads to a cut-off of the cosmic ray spectrum below 1020 eV. The charged mesons generated in these interactions initiate a decay chain that results in neutrinos. Since the mesons and muons do not lose energy before decay, the high energy end of the spectrum of these neutrinos follows the injection spectrum of UHECR, while below the interaction threshold it is flat [4], [5]. The neutri- nos which are produced by photomeson producing inter- actions of UHECR nuclei are sometimes referred to as cosmogenic neutrinos. Several calculations of of the fluxes of UHE photome- son neutrinos were published in the 1970s [3, 4, 5, 6, 7], Hill and Schramm [8, 9] used the non-detection of such neutrinos to place an upper limit on the cosmological evo- lution of the sources of UHECR. The problem has been revisited several more times [10, 11, 12]. In 2004 Stanev [13] considered interactions of UHECR with photons of the extragalactic infrared and optical background (IRB), pointing out that this process gener- ates non-negligible cosmogenic neutrino fluxes. This sug- gestion was quickly followed by a confirmation in Ref. [14] which emphasized the importance of the IRB as interac- ∗Electronic address: [email protected] †Electronic address: [email protected] ‡Electronic address: [email protected] tion target. This idea was further developed in Ref. [15]. Ref. [13] gave an estimate of the cosmogenic neutrino flux generated in interactions on the IRB, but did not account correctly for the cosmological evolution of the infrared background. In this paper we perform a cal- culation using a realistic empirically based model of the cosmological evolution of the spectral energy distribution of the extragalactic IR-UV background given in Ref. [16] which will be referred to as SMS05. The aim is to es- timate correctly the role of these extragalactic photons, particularly the infrared photons which are by far the most numerous, as targets for UHE proton interactions. in Section II we discuss the model of the infrared background and its cos- mological evolution. In Section III we describe the cal- culation. Section IV gives the results of the calculation and Section V contains the discussion of the results and the conclusions from this research. The paper is organized as follows: II. COSMOLOGICAL EVOLUTION OF THE IR-UV BACKGROUND It is now well known that galaxies had a brighter past owing to the higher rate of star formation which took place. Strong evolution is supported by many observa- tions relating IR luminosity to the much higher star for- mation rate at z ∼ 1 and to the recent determination that most Lyman break galaxies at z ∼ 1 are also luminous infrared galaxies. In addition to the evolution of galaxy luminosity, some increase in galaxy number density is ex- pected owing to the hierarchical clustering predicted by cold dark matter models. However, luminosity evolution is the dominant effect and it is difficult to separate out a component of density evolution. In order to calculate intergalactic IR photon fluxes and densities and their evolution over time (or redshift), SMS05 performed an empirically based calculation the SED of the IBR (infrared background radiation) by using (1) the luminosity dependent galaxy spectral energy dis- tributions (SEDs) based on galaxy observations, (2) ob- servationally derived galaxy luminosity distribution func- tions (LFs) and (3) the latest redshift dependent lumi- nosity evolution functions, sometimes referred to as Lilly- Madau plots. The SMS05 calculation was an improved version of the work presented in Refs. [17], [18] and [19]. The calculation considers two different cosmological evolutions, E(z) baseline and fast, of the infrared emission of the type 3 - m c / , ε d n d ε E(z) =   : z < zflat (1 + z)m (1 + zflat)m : zflat < z < 6 0 : z > 6 (1) The baseline evolution model is described by m=3.1 and zflat = 1.3, while the fast evolution model uses m=4 and zflat = 1. The infrared emission at z > zflat is constant in both models. Figure 1 shows the number density between 101 100 10-1 10-2 10-3 2 z=0 z=1 z=2 z=3 z=4 z=5 10-2 10-1 ε, eV 100 FIG. 2: Number density of the infrared background at dif- ferent redshifts calculated by SMS05 [16] in the fast evolution model. basline fast III. THE CALCULATION 3 - m c , ε N 4 3.5 3 2.5 2 1.5 1 0.5 0 0 1 2 3 4 5 redshift FIG. 1: Number density of the IRB at different redshifts as calculated by SMS05 [16]. photon energy of 3.16×10−3 and 1 eV in both models. The fast evolution model has higher density in the current cosmological epoch as well at the IRB maximum epoch, which is around z = 2. The increase of the total IRB number density increases by a factor of about 4 from z = 0 to z = 2 and decreases at higher redshifts. One should note, however, that the cosmological evolution of the infrared background density is much slower than that of MBR since the current IRB density is accumulated from the infrared emission of different sources since z = 6. Figure 2 shows the energy spectrum of the fast infrared background at redshifts from 0 to 5. One can see both the increase of the total photon density as well as the shift of the maximum of the emission to higher energy at higher redshifts. In terms of photoproduction interactions on IRB this means that lower energy cosmic rays will be above the photoproduction threshold at higher redshifts. Both figures above show the number density of IRB rather than the usual presentation of the energy density. Since we are using the infrared background as a target for cosmic ray interactions this is the relevant quantity. The calculation was performed in two stages: (1) cal- culation of the neutrino yields from interactions with ex- tragalactic infrared photons and (2) a subsequent inte- gration of the yields to obtain the cosmogenic neutrino flux from such interactions. This approach gives us the flexibility to easily obtain the neutrino flux using dif- ferent parametrizations of the cosmic ray emissivity, in- jection spectra and cosmological evolution of the cosmic ray sources. This approach, however, suffers from the fact that since the yields are calculated only on the IRB, they do not account for the fact that high energy protons interact mainly with the much more numerous MBR pho- tons. A. Calculation of the Neutrino Yields The neutrino yields as a function of the proton energy Ep, neutrino energy Eν and the redshift, z, were cal- culated using the IRB spectra at different cosmological epochs, i.e., as a function of redshift, that were provided by the authors of Ref. [16]. Each of the yield calcula- tions was performed for proper distances corresponding to ∆z=0.2 using an ΩM =0.3, ΩΛ=0.7 cosmology as D(z) = c H0 Z zmax zmin 1 1 + z (cid:2)ΩM (1 + z)3 + ΩΛ(cid:3)−1/2 (2) The IRB is considered to be constant during each cos- mological epoch of duration ∆z=0.2. In this way the matrix element corresponding to dt/dz dependence was accounted for in the yields. The yields were calculated with the code used in Ref. [12] and the photoproduction interaction code SOPHIA [20]. All gen- erated neutrinos are redshifted by the code to the end of the ∆z epoch. The yields were calculated for redshifts 0 < z < 5 and for cosmic ray energies above 1018 eV in ten logarithmic bins per decade of energy. 10-2 10-3 10-4 10-5 10-7 1 - c p M l , ν E n d / ν N d mbr irb 1020 1019 1018 10-8 1014 1015 1016 1017 Eν, eV 1018 1019 1020 FIG. 3: Neutrino yields for 1020 eV protons interacting with both MBR photons and for 1018 eV, 1019 eV and 1020 eV protons interacting with IRB photons at z = 0, given for protons traveling a distance of 1 Mpc. Fig. 3 compares the νµ yields for UHE protons travel- ing a distance of 1 Mpc and interacting with MBR and IRB photons. The yield for 1020 eV protons interacting with IRB photons is about a factor of 10 lower than that for MBR interactions. This difference is much smaller than the ratio of the MBR and IRB total densities, and demonstrates that 1020 eV protons interact mainly with photons in the higher frequency Wien tail of the 2.7K MBR spectrum. Protons of energy 1019 eV do not interact at z=0 with MBR photons, but they readily interact and produce neutrinos by interactions with IRB photons, as do pro- tons of energy 1018 eV. Even protons of energy 1017 eV occasionally interact with IRB photons, but their con- tribution is very small and is neglected in this calcula- tion. Even for E−2 cosmic ray spectra, the smaller 1019 and 1018 eV yields are multiplied by the much higher flux of cosmic rays with such energies. This is the basis of the significant neutrino (and γ-ray) production from UHECR-IRB interactions. B. Integration of the Yields 3 shift evolution of flat spectrum radio sources, given as an analytic approximation in Ref. [21]. We use the fast evo- lution model from SMS05 since it is more consistent with the new observations of the Spitzer telescope [22, 23]. We normalize the cosmic ray energy flux at Ep = 1019 eV to EpdNp/Ep = 2.5×10−18 cm−2s−1sr−1 [24, 25]. Since the calculation is extended to energies below 1019 eV, the code uses the cosmic ray flux at 1019 eV to cal- culate the injection spectra at lower and higher energy. Therefore, the cosmic ray emissivity above 1018 eV de- pends on the injection spectrum. The injection spectrum itself is used as a free parameter in order to study its in- fluence on the cosmogenic neutrino spectrum. The integration procedure also has to account for the modification of the cosmic ray spectrum owing to inter- actions with MBR photons. This was done in two crude, but reasonable, ways. The first one is the introduction of a high energy cutoff of the spectrum as a function of the redshift. The second one, which is used in the results presented below, is to weight the cosmic ray in- jection spectrum with the interaction length λIRB on the infrared background radiation. The cosmic rays interact- ing in the IBR used in the integration of the yields are FCR λIRB/λtot, where λtot is the interaction length in the total IRB and MBR fields. The fraction of the cos- mic ray flux used in the integration procedure is shown in Fig. 4. If one arbitrarily determines the high energy B R I λ λ / h P 100 10-1 10-2 10-3 10-4 1018 0 1 2 3 4 5 1019 Ep, eV 1020 1021 The second phase of the calculation requires the parametrization of the redshift evolution of the emis- sivity of cosmic ray sources, and the form of the cos- mic ray injection spectrum. We assume a cosmic ray injection spectrum of the power-law form dN/dEp = AE−(γ+1) exp(−Ep/Emax) with Emax = 1021.5 eV. p We consider here two empirically based models for the evolution of UHECR power with redshift, viz., (1) one based on the redshift evolution of the star formation rate that was used in the calculation of the infrared back- ground in SMS05, and (2) the other based on the red- FIG. 4: Fraction of the total cosmic ray flux used in the in- tegration of the neutrino yields from interactions in the IRB. The different lines correspond to fractions at different red- shifts as indicated by the numbers in the plot. cutoff of the cosmic ray energy spectrum as the energy at which only 10 per cent of the cosmic rays interact in the IRB, it would be 7×1019 eV at z = 0 compared to 2.5×1019 and 4×1018 eV at z = 1 and 5. Since the yields include the dt/dz factor the integration becomes very simple, viz.: dN i ν/dEν = Z 5 ×Z Ec dNp dEp dz × 0 E(z) Y i [(1 + z)Eν; Ep, z] dEp , (3) where the index i indicates the neutrino flavor. IV. RESULTS The results of the integration are shown in Fig. 5. The top panel of the figures compares the fluxes of cosmo- genic νµ + ¯νµ neutrinos generated by interactions with MBR photons (histogram) with those generated by in- teractions with IRB photons (squares), assuming a γ = 1 UHECR injection spectrum with the fast evolution of the emissivity of the cosmic ray sources. All panels of Fig. 5 are calculated with the same cosmological evolution. Us- ing the baseline evolution model will give neutrino fluxes which are about 25-30% lower. The peak flux of the IRB-generated neutrinos is not much lower than that of the MBR-generated ones, i.e., 1.7×10−17 compared to 2.2×10−17 cm−2s−1sr−1. The peak is, however, shifted to lower Eν by about a factor of 3. The main reason for that shift is the contribution of protons of energy below 3×1019 eV to the neutrino production. The IRB-generated neutrino flux is also de- pleted at energies above 1019 eV. This is because protons of energy above 5×1019 eV very rarely interact with IRB photons before they lose their energy in MBR interac- tions. At energies below the peak the IRB-generated neutrino flux is somewhat flatter than the MBR one, al- though the statistical uncertainty of the calculation does not allow us to make a quantitative statement regarding this. The middle panel of the figure shows the IRB- generated fluxes of νe's and ¯νe's assuming an E−2 in- jection spectrum. The electron neutrino flux peaks at the same energy as the muon neutrino one. The ¯νe flux, which is due mostly to neutron decay neutrinos, is shifted and widened at its lower energy end. The dip between the νe and ¯νe peaks is not as deep as it is in the MBR neutrino case. The reason for that is that the νe peak is somewhat wider at energies below the peak. The bottom panel of Fig. 5 shows the fluxes of IRB-generated νµ + ¯νµ assuming a steeper injection spectrum γ = 1.5 and m = 3.1. There are two main differences from the γ=1 case. The peak of the IRB-generated neutrino flux is higher by almost a factor of 3 (6.5×10−17 in the same units) than for the MBR-generated neutrinos and this peak and is shifted down in energy by a factor of ∼ 3 to ∼ 1016.5 eV. This is caused by the higher flux of protons of energy below the high energy cutoff. In the case of the MBR- generated neutrinos, the general effect is not as strong but is reversed; the steeper proton spectrum results in a lower flux. 4 γ = 1.0 IRB MBR 10-15 10-16 10-17 10-18 10-19 10-20 1012 1013 1014 1015 1016 1017 Eν, eV 1018 1019 1020 1021 γ = 1.0 IRB 10-15 10-16 10-17 10-18 10-19 10-20 1012 1013 1014 1015 1016 1017 Eν, eV 1018 1019 1020 1021 γ = 1.5 IRB 10-15 10-16 10-17 10-18 10-19 1 - r s 1 - s 2 - m c , ν E n d / ν N d l 1 - r s 1 - s 2 - m c l , ν E n d / ν N d 1 - r s 1 - s 2 - m c l , ν E n d / ν N d 10-20 1012 1013 1014 1015 1016 1017 Eν, eV 1018 1019 1020 1021 FIG. 5: Top panel: νµ (solid) and ¯νµ (dashed) spectra gen- erated by interactions with IRB photons. Their sum (open squares) is compared to those generated by interactions with MBR (MBR) photons (full squares) for γ=1 assuming fast evolution of the cosmic ray source emissivity. Middle panel: νe (solid) and ¯νe (dash) spectra for injection spectra as in the top panel. Their sum is shown with open squares. Bottom panel: νµ + ¯νµ spectra for injection spectrum with γ=1.5. Because of the very strong dependence of the flux of cosmogenic neutrinos on the cosmological evolution of the cosmic ray sources [26], we investigated this depen- dence further. Fig. 6 compares the baseline and fast cos- mological evolutions of Ref. [16] to these of Refs. [21, 24]. All evolution models are normalized to 1 at present, i.e. for z = 0. n o i t u o v e l l i l a c g o o m s o c 25 20 15 10 5 0 normalized to 1 at z=0 baseline fast W95 DP1990 5 γ = 1.0 fast baseline m=3 D&P 10-15 10-16 10-17 10-18 10-19 1 - r s 1 - s 2 - m c l , ν E n d / ν N d 0 1 2 3 4 5 redshift 10-20 1012 1013 1014 1015 1016 1017 Eν, eV 1018 1019 1020 1021 FIG. 6: Four different models for the cosmological evolu- tion of the cosmic ray sources (see text for description and references). FIG. 7: Cosmogenic neutrino fluxes for four different models of the cosmological evolution of the cosmic ray sources - see text. The evolution taken from Ref. [24], that was used for calculations of the cosmogenic neutrino flux from interac- tions in the MBR [12], gives a UHECR emissivity which is about 60% higher at redshift of 2 than the average of the models of SMS05. The cosmological evolution of the flat spectrum radio galaxies [21] has an intermediate red- shift evolution; it is faster than m = 3 and slower than m = 4 below z = 1 and peaks at about z = 2. It is also distinguished by its rapid decrease in emissivity at z > 3. Figure 7 compares the cosmogenic neutrino fluxes of νµ + ¯νµ generated by the baseline and fast models of SMS05 with the m = 3 model used in Ref. [12] and that of Ref. [21]. The difference in the calculated fluxes is ac- tually quite small, compared to all other uncertainties of the calculation. The fast and the baseline models bracket from above and from below the fluxes of cosmogenic neu- trinos from interactions in the IRB, while the other two models fall between these two. The main reason for the small differences is the dz/dt matrix element that de- creases the contribution of higher redshifts because in the cosmological integration the emissivity is multiplied by the smaller time intervals involved at higher redshifts. In the case where UHECR luminosity evolution is as- sumed to be proportional to the redshift distribution of flat spectrum radio galaxies as given in Ref. [21], the neu- trino spectra peak at a somewhat higher energy because of the relatively small UHECR emissivity at the higher redshifts. In Fig. 8 we present the total fluxes of cosmogenic neu- trinos from interactions in the MBR and IRB for injection spectra with γ = 1.0 and 1.5 and for fast cosmological evolution of the cosmic ray sources as in Ref. [16]. For a relatively flat (γ = 1) injection spectrum (empty squares in Fig. 8) interactions with IRB photons gener- ate almost as many cosmogenic neutrinos as interactions on MBR. The peak of the total neutrino energy spec- trum from interactions in the MBR and in IRB is shifted to lower energy by a small amount (from 3×1017 eV to fast evolution IRB MBR γ=1.0 γ=1.5 10-15 10-16 10-17 10-18 10-19 1 - r s 1 - s 2 - m c l , ν E n d / ν N d 10-20 1012 1013 1014 1015 1016 1017 Eν, eV 1018 1019 1020 1021 FIG. 8: Total muon νµ + ¯νµ fluxes for γ = 1.0 (empty squares) and 1.5 (full squares) calculated with fast cosmic ray source evolution. The shaded area represents the W&B [27] upper bound on astrophysical source neutrinos. about 2×1017 eV). The distribution extends to lower neu- trino energies by more than half a decade. For steeper spectra (γ = 1.5) the contribution of IRB- generated neutrinos is more significant and the result- ing flux is almost an order of magnitude larger. The magnitude of the flux at the peak of the spectrum is ∼ 10−16 cm−3s−1sr−1 and is higher than that of the MBR-generated neutrinos by a factor of ∼ 3. In both cases there is no increase of the neutrino flux at energies exceeding 1019 eV. The influence of increased cosmological evolution of the cosmic ray sources is practi- cally the same as in the case of MBR-generated neutrinos alone. In this context, we note that the cosmological evo- lution of target IR photon density is slower that of the MBR. For comparison, the shaded area in Fig. 8 shows the upper bound on the astrophysical neutrino spectra given in Ref. [27]. The lower edge is the bound in absence of cosmic ray source cosmological evolution, and the upper TABLE I: Rates per km3 water per year of showers above different energy generated by different types of neutrino in- teractions for cosmic ray power density P0 at z=0 of 1.4×1031 W Mpc−3 and fast cosmological evolution for homogeneously distributed cosmic ray sources (see text). γ=1 γ=1.5 log10 Esh (GeV) > MBR IRB MBR IRB 6 0.092 0.021 0.078 0.085 7 0.088 0.019 0.072 0.072 8 0.079 0.010 0.063 0.030 9 0.044 0.001 0.027 0.001 10 0.008 0.000 0.003 0.000 edge is for (1 + z)3 evolution. The inclusion of the proton-IRB interactions somewhat reverses the trend of the injection spectrum dependence of the cosmogenic neutrino flux. Without including such interactions, steeper injection spectra lead to smaller cos- mogenic neutrino fluxes; with the inclusion of the contri- bution to the neutrino flux from interactions of IRB pho- tons with relatively lower energy protons, steeper cosmic ray spectra generate higher neutrino fluxes. The reason is that we normalize the cosmic ray injection spectrum at 1019 eV, which is now in the middle of the energy range of the interacting cosmic rays. One can see in Fig. 4 the dominance of the interactions in the MBR of cosmic rays of energy above 1019 at all redshifts higher than 1. For steeper cosmic ray injection spectra the number of such particles is decreased while that of cosmic rays below 1019 eV, that interact in the IRB, is significantly increased. Because of the lower average energy of the IRB- generated neutrinos, the spectra are shifted and their de- tectability is lower than that of MBR generated photons. This is because the neutrino-nucleon cross section rises monotonically with energy. Table I shows the shower rates of νe and ¯νe CC (charged current) interactions per km3yr of water detector for cosmogenic neutrinos gener- ated by interactions with MBR and IRB photons. These rates are the products of the neutrino cross section times the neutrino flux integrated above Esh. We assume that the total neutrino energy is transfered to the shower ini- tiated by its CC interaction. The calculation of event rates of CC and NC interactions of muon and tau neu- trinos are much more difficult and require Monte Carlo models of particular experiments. The second column of Table I corresponds to the num- bers given in a similar table in Ref. [12]. The numbers can not be directly compared because of the different cos- mologies (ΩM = 1 in Ref. [12] and ΩM = 0.3 here) and cosmological evolutions of the cosmic ray sources used. The ΩM = 0.3 cosmology increases the neutrino rates by about 70%. In the γ = 1.0 injection case the rates from MBR neu- trinos are higher by factors above 4 at all shower thresh- olds, while in the γ = 1.5 case the IRB rates are higher or similar for shower thresholds below 108 GeV. Above that energy MBR neutrinos generate higher rate. Note 6 γ=1.7 D&P m=0 10-15 10-16 10-17 10-18 10-19 1 - r s 1 - s 2 - m c l , ν E n d / ν N d 10-20 1012 1013 1014 1015 1016 1017 Eν, eV 1018 1019 1020 1021 FIG. 9: Total muon νµ + ¯νµ fluxes for γ = 1.7 and evolution from Ref. [21] source evolution (full squares) and no evolution (empty squares). that IRB neutrinos do not contribute at all to the shower rates above 109 GeV. The total shower rate for Esh > 106 GeV is higher than the MBR only rate by about 20% in the γ=1.0 case, while in the γ = 1.5 case it more than doubles the detection rate. Most of the contemporary fits of the injection spectrum of the highest energy cosmic rays confirm the conclusion of Ref. [28] that it is steeper than E−2. The spectrum derived by these authors is E−2.7 with a significant flat- tening at about 1018 eV, which could be explained with different effects, see e.g. Ref. [29]. The shape of the spec- trum may be accounted for in this model as a result of the pγ → e+e− process [30] as discussed in Ref. [31]. This pair production process creates a dip at about 1019 eV and a slight excess at the transition from pair production to purely adiabatic proton energy loss at about 1018 eV. This fit does not require a strong cosmological evolution of the cosmic ray sources, but can accommodate a mild one ∝ (1 + z)m with m ≤ 3 [32]. In the case of flatter injection spectrum the pair production dip is well fit also by m = 4. Figure 9 shows the spectra of the cosmogenic neutrinos from interactions with MBR and IRB photons assuming a steep injection spectrum with γ = 1.7 and (1) no evolution with (m = 0) and (2) evolution according to Ref. [21]. The difference between the two neutrino spec- tra is significant; the peak values are 10−17(1.5×10−16) cm−2s−1sr−1 for without and with evolution. The ad- dition of the IRB component brings these spectra into the range of detectability, especially in the case of mild cosmological evolution. V. DISCUSSION AND CONCLUSIONS The evolution models of SMS05 do not give the highest IRB-generated neutrino flux. We compared the IRB pho- ton density in this model with the models of Refs. [33, 34]. 7 Both of these models have higher IRB density in the rel- evant energy range between 3×10−3 and 1 eV. The total IRB densities in the this range are 1.27 [33] and 1.12 [34] compared with the density of 1.03 used here. In addi- tion, Ref. [34] shows somewhat faster cosmological evo- lution. The use of any of these models would have in- creased somewhat the calculated flux of cosmogenic neu- trinos. The uncertainty in the IRB flux is of the order of 30% [16], while the biggest uncertainty in this calculation is in the UHECR flux and its cosmological evolution. The IRB contribution to the total cosmogenic neutri- nos flux can be slightly increased as protons of energy below 1018 eV can interact with IRB photons and gener- ate lower energy neutrinos. If such interactions were in- cluded the IRB spectrum would be wider than the MBR one, especially at energies below 1016 eV. It is difficult to compare our results with those of Refs. [14, 15] because of the different astrophysical input in these calculations. Qualitatively the results of these calculations are similar to ours and certainly agree within a factor of 2. In conclusion, we calculated the flux of cosmogenic neutrinos from interactions of UHECR protons with IRB photons using the recent calculations of IR photon spec- tra densities as a function of redshift by SMS05. Our calculations show that UHECR interactions with IRB photons produce a significant flux of cosmogenic neutri- nos, one which is comparable to interactions with MBR photons. This is especially true in the case of assumed steep injection spectra of the ultrahigh energy cosmic rays. The total neutrino event rates at energies above 1 PeV increase by more than a factor of 2 in the case of injection spectra with γ = 1.5. Because of the much lower mean free path of protons above 1020 eV in the MBR interactions with IRB photons do not increase the higher energy end of the cosmogenic neutrino spectrum. The total cosmogenic fluxes, however, are still not de- tectable with conventional neutrino telescopes such as IceCube [35] or the European km3 telescope [36]. A reli- able detection is only expected from radio [37] and acous- tic neutrino detectors. Acknowledgments We thank Tanja Kneiske for shar- ing with us the tables of cosmological evolution of the IR background from Ref. [34]. This work was supported in part by NASA grants ATP03-0000-0057 and ATP03- 0000-0080. [1] V.S. Berezinsky and G.T. Zatsepin, Phys. Lett. 28b, 423 (2001). (1969); Sov. J. Nucl. Phys. 11, 111 (1970). [2] K. Greisen, Phys. Rev. Lett. 16, 748 (1966); G.T. Zat- sepin and V.A. Kuzmin, JETP Lett. 4, 78 (1966). [3] J. Wdowczyk, W. Tkaczyk and A.W. Wolfendale, J. Phys. A 5, 1419 (1972). [4] F.W. Stecker, Astroph. Space Sci. 20, 47 (1973). [5] F.W. Stecker, Astrophys. J. 238, 919 (1979). [6] V.S. Berezinsky and A.Yu. Smirnov, Astroph. Space Sci. [20] A. Mucke, R. Engel, J.P. Rachen, R.J. Protheroe and T. Stanev, Comput. Phys. Commun. 124, 290 (2000) M. Murgia et al., preprint: astro-ph/0406225 [21] J.S. Dunlop and J.A. Peacock, MNRAS 247, 19 (1990). [22] P.G. P´erez-Gonzalez et al., ApJ, 630, 82 (2005). [23] E. Le Floc'h et al., ApJ, 632, 169L (2005). [24] E. Waxman, Astrophys. J. 452, L1 (1995). [25] F.W. Stecker and S.T. Scully, Astropart. Phys. 23, 203 32, 461 (1975). (2005) [7] V.S. Berezinsky and G.T. Zatsepin, Sov. J. Uspekhi, 20, [26] D. Seckel and T. Stanev, Phys. Rev. Lett.,95, 141101 361 (1977) (2005) [8] C.T. Hill and D.N. Schramm, Phys. Rev. D31, 564 [27] E. Waxman E and J.N. Bahcall, Phys. Rev. D59:023002 (1985). (1999) [9] C.T. Hill, D.N. Schramm and T.P. Walker, Phys. Rev. D34, 1622 (1986). [10] S. Yoshida and M. Teshima M., Progr. Theor. Physics 89, 833 (1993). [28] V.S. Berezinsky, A.Z. Gazizov and S.I. Grigorieva, Phys. Lett. B612, 147 (2005); see also hep-ph/0204357 and astro-ph/0210095. [29] R. Aloisio and V.S. Berezinsky, Astrophys. J. 625, 249 [11] R.J. Protheroe and P.A. Johnson, Astropart. Phys. 4, (2005). 253 (1996). [12] R. Engel, D. Seckel and T. Stanev, PRD 64, 093010 (2001) [13] T. Stanev, Phys. Lett. B595 50 (2004)] [14] E.V. Bugaev, A. Misaki and K. Mitsui, Astropart. Phys., [30] G.R. Blumenthal, Phys. Rev.D1, 1596 (1970) [31] V.S. Berezinsky and S.I. Grigorieva, Astron.and Astro- phys. 199, 1 (1988). [32] D. De Marco and T. Stanev, Phys. Rev. D72, 081301 (2005). 24:345. (2005) [33] A. Franceschini et al., Astron.and Astrophys. 378, 1 [15] E.V. Bugaev and P.A. Klimai, astro-ph/0509395 [16] F.W. Stecker, M.A. Malkan and S.T. Scully, Ap. J., sub- (2001) [34] T.M. Kneiske, K. Mannheim and D.H. Hartman, As- mitted, astro-ph/0510449 (2005). tron.and Astrophys. 386, 1 (2002). [17] M.A. Malkan and F.W. Stecker, Astrophys. J. 496, 13 (1998). [18] M.H. Salamon and F.W. Stecker, Astrophys. J. 493, 547 (1998). [19] M.A. Malkan and F.W. Stecker, Astrophys. J. 555, 641 [35] http://icecube.wisc.edu [36] http://www.km3net.org [37] H. Falke, P. Gorham and R.J. Protheroe, New Astron. Rev. 48:1487 (2004)
0708.2440
1
0708
2007-08-17T21:38:49
Spitzer Observations of Gamma-Ray Burst Host Galaxies: A Unique Window into High Redshift Chemical Evolution and Star-formation
[ "astro-ph" ]
We present deep Spitzer 3.6 micron observations of three z~5 GRB host galaxies. Our observations reveal that z~5 GRB hosts are a factor of 3 less luminous than the median rest-frame V-band luminosity of spectroscopically confirmed z~5 galaxies in the GOODS fields and the UDF. The strong connection between GRBs and massive star formation implies that not all star-forming galaxies at these redshifts are currently being accounted for in deep surveys and GRBs provide a unique way to measure the contribution to the star-formation rate density from galaxies at the faint end of the galaxy luminosity function. By correlating the co-moving star-formation rate density with co-moving GRB rates at lower redshifts, we estimate a lower limit to the star-formation rate density of 0.12+/-0.09 and 0.09+/-0.05 M_sun/yr/Mpc^3 at z~4.5 and z~6, respectively. Finally, we provide evidence that the average metallicity of star-forming galaxies evolves as (stellar mass density)^(0.69+/-0.17) between $z\sim5$ and $z\sim0$, probably indicative of the loss of a significant fraction of metals to the intergalactic medium, particularly in low-mass galaxies.
astro-ph
astro-ph
Submitted to The Astrophysical Journal Spitzer Observations of Gamma-Ray Burst Host Galaxies: A Unique Window into High Redshift Chemical Evolution and Star-formation R. Chary1, E. Berger2,3,4, L. Cowie5 ABSTRACT We present deep 3.6 µm observations of three z ∼ 5 GRB host galaxies with the Spitzer Space Telescope. The host of GRB 060510B, at z = 4.942, is detected with a flux density of 0.23 ± 0.04 µJy, corresponding to a rest-frame V −band luminosity of 1.3 × 1010 L⊙, or ≈ 0.15 L∗,V,z=3. We do not detect the hosts of GRBs 060223A and 060522 and constrain their rest-frame V −band luminosity to < 0.1 L∗,V,z=3. Our observations reveal that z ∼ 5 GRB host galaxies are a factor of ∼ 3 less luminous than the median luminosity of spectroscopically-confirmed z ∼ 5 galaxies in the Great Observatories Origins Deep Survey (GOODS) and the Hubble Ultra Deep Field (UDF). The strong connection between GRBs and massive star formation implies that not all star-forming galaxies at these redshifts are currently being accounted for in deep surveys and GRBs provide a unique way to measure the contribution to the star-formation rate density from galaxies at the faint end of the galaxy luminosity function. By correlating the co-moving star-formation rate density with co-moving GRB rates at lower redshifts, we estimate a lower limit to the star-formation rate density of 0.12 ± 0.09 and 0.09 ± 0.05 M⊙ yr−1 Mpc−3 at z ∼ 4.5 and z ∼ 6, respectively. This is in excellent agreement with extinction corrected estimates from Lyman-break galaxy samples. Finally, our observations provide initial evidence that the metallicity of star- forming galaxies evolve more slowly than the stellar mass density between z ∼ 5 and z ∼ 0, probably indicative of the loss of a significant fraction of metals to the intergalactic medium, especially in low-mass galaxies. 1Spitzer Science Center, California Institute of Technology, Mail Stop 220-6, Pasadena, CA 91125 2Observatories of the Carnegie Institution of Washington, 813 Santa Barbara Street, Pasadena, CA 91101 3Princeton University Observatory, Peyton Hall, Ivy Lane, Princeton, NJ 08544 4Hubble Fellow 5Institute for Astronomy, University of Hawaii, HI 96822 -- 2 -- Subject headings: gamma rays:bursts -- cosmology:observations -- galaxies:high- redshift -- galaxies:starburst -- galaxies:abundances 1. Introduction Our ability to measure the star-formation rate density (SFRD) at z > 4 relies al- most entirely on either narrow-band surveys which detect strong Lyα emitting galaxies (Hu et al. 2004; Nagao et al. 2007) or deep imaging surveys of UV bright Lyman-break galaxies (Giavalisco et al. 2004; Bouwens et al. 2006; Hu & Cowie 2006). These surveys, by virtue of being flux limited, trace the bright end of the galaxy luminosity function down to ∼ 0.04 L∗,UV,z=3. The various observations have revealed a decline in the SFRD by about a factor of 3 between z ∼ 3 and z ∼ 6 (Bouwens et al. 2006; Bunker et al. 2004), with much of this decline being due to the evolution of the bright end of the galaxy luminosity function. More than 90% of the estimated SFRD at these redshifts takes place in sub-L∗,UV,z=3 galaxies. In addition, spectroscopic confirmation of high redshift galaxies relies on Lyα emission, which is easily obscured by dust. There is now increasing evidence for rapid dust production within ∼ 1 Gyr of the Big Bang (Chary et al. 2007, 2005; Maiolino et al. 2004). As a result, quantifying possible dust extinction corrections and measuring the faint-end slope of the galaxy luminosity function is essential for minimizing uncertainties in the high redshift SFRD. Measurements of the metallicity in typical star-forming galaxies at z > 4 is beyond the technological capability of the current generation of instrumentation. The relevant rest-frame optical emission lines are very weak and are redshifted to the mid-IR. The al- ternative approach of studying chemical enrichment through damped Lyα absorbers (DLAs) detected against background quasars appears to be limited to z . 5 (Prochaska et al. 2003; Songaila & Cowie 2002), and is biased towards tracing the properties of extended halo gas, which at lower redshift, significantly underestimates the disk metallicity. As a result, the ap- parent evolution in the mass-metallicity and luminosity-metallicity relations (M-Z and L-Z) from z = 0 to z ∼ 2 (e.g. Tremonti et al. 2004; Kobulnicky & Kewley 2004; Savaglio et al. 2005) cannot be traced to z > 5, where such information should shed light on the initial stages of mass build-up and metal enrichment in galaxies. Long duration GRBs, by virtue of being associated with the deaths of massive stars, provide a complementary technique for measuring the SFRD and the chemical enrichment history. Swift has revolutionized this study by detecting GRBs out to z ∼ 6 (Gehrels et al. 2004; Kawai et al. 2006). Prompt spectroscopy of the bright afterglows has now provided -- 3 -- a sample of ∼ 20 GRBs over a wide range of redshifts with a wealth of metal absorption features arising in the host galaxy (e.g Jensen et al. 2001; Castro et al. 2003; Hjorth et al. 2003; Vreeswijk et al. 2004; Fynbo et al. 2006). These observations provide a unique win- dow into the metallicity and gas column density in star forming environments at high redshifts (Chen et al. 2005; Berger et al. 2006b; Prochaska et al. 2006; Price et al. 2007; Prochaska et al. 2007a). Once the afterglows fade away, deep observations of the field can also reveal the stellar mass and star-formation rate of the host galaxies, which can then be correlated with the inferred metallicities. In order to study the host galaxies of high redshift GRBs and take advantage of the diagnostics afforded by GRBs, we present Spitzer Space Telescope 3.6 µm observations of the hosts of three GRBs at 4 . z . 5. Building on the constraints provided by Berger et al. (2006a) on the host galaxy of GRB 050904 at z = 6.295, we discuss the nature of the host galaxies, redshift evolution of the luminosity-metallicity relation and provide an independent measure of the high redshift SFRD for comparison with estimates from Lyman-break galaxy samples. Throughout this paper, we adopt a ΩM = 0.27, ΩΛ = 0.73, H0=71 km s−1 Mpc−1 cosmology. 2. Observations As part of Spitzer program GO 20000 (PI: Berger) we observed the fields of GRBs 060223A (z = 4.406; Berger et al., in prep.), 060510B (z = 4.942; Price et al. 2007), and 060522 (z = 5.110; Berger et al., in prep.) with the Infrared Array Camera (IRAC; Fazio et al. 2004) in the bandpasses centered at 3.6 and 5.8 µm (Table 1). The observations were undertaken between September and November 2006, after the afterglows associated with the GRBs had faded below the detectability threshold. As shown in Table 1, the GRB fields lie in regions with "low" to "medium"-level zodiacal background and cirrus of 13−28 MJy/sr at 24 µm on the date of the observations. We used 100 s integrations with about 130 medium scale dithers from the random cycling pattern for total on-source integration times of ∼ 13000 s at each passband. The nominal 3σ point source sensitivity limits are 0.26 and 2.4 µJy, respectively. Starting with the S14.4.0 pipeline-processed basic calibrated data (BCD) sets we cor- rected the individual frames for muxbleed and column pull down using software developed for the Great Observatories Origins Deep Survey (GOODS). Due to the presence of bright stars in the field, many of the frames at 3.6 µm also showed evidence for "muxstriping". This was removed using an additive correction on a column by column basis. The processed BCD frames were then mosaiced together using the MOPEX routine (Makovoz & Khan 2005) and -- 4 -- drizzled onto a 0.6′′ grid. Astrometry was performed with respect to the brightest 2MASS stars in the field which showed a peak-to-peak astrometric uncertainty of 0.2′′ at 3.6 µm. The location of the GRB hosts was determined by aligning the Spitzer images against images of the afterglow from the Swift UV/optical telescope (060223A and 060522) and the Gemini Multi-Object Spectrograph on the Gemini-north 8-m telescope (060510B). For the latter, the astrometric uncertainty is 0.09′′ in each coordinate, while using the UVOT images we obtain an astrometric uncertainty of about 0.6′′. All three GRB locations show the presence of nearby (3′′) brighter galaxies: GRB 060223A has two sources with flux densities of 7.2 and 9.1 µJy at distances of 1.9′′ and 2.3′′ from the GRB position; GRB 060510B has a source with a flux of 6.1 µJy about 3.1′′ from the GRB po- sition; and GRB 060522 has a source with 0.72 µJy located 1.6′′ away from the burst position. This is not unexpected given the high source density in deep IRAC images. We subtracted the contribution of these sources, in order to obtain the strongest possible constraints on the flux from the GRB host galaxies. Photometry at the position of the host galaxies was performed in fixed circular aper- tures of 1.2′′ radius with appropriate beam size corrections applied as stated in the Spitzer Observer's Manual. We clearly detect a galaxy coincident with the position of GRB 060510B with a flux density of 0.23 ± 0.04 µJy at 3.6 µm, and a 3σ upper limit of 2.4 µJy at 5.8 µm (Figure 1). For GRBs 060223A and 060522, due to blending from nearby brighter sources and the residual effects from muxbleed, we are only able to provide 3σ upper limits to the flux of the host galaxy (Table 1). 3. Luminosity and Metallicity of GRB Hosts Of the 4 GRB host galaxies at z ∼ 5 observed in this program at 3.6 and 5.8 µm, (including GRB 050904), only GRB 060510B is clearly detected. The observed 3.6 µm flux densities/limits for these galaxies correspond to rest-frame V −band luminosities of ∼ 0.15 L∗,V,z=3, where L∗,V,z=3 is about 8 × 1010 L⊙ (Marchesini et al. 2007; Shapley et al. 2001). It is illustrative to compare the properties of GRB hosts with the field galaxy population at similar redshifts. The GOODS fields have extensive spectroscopy of galaxies at high redshift (Vanzella et al. 2005; Vanzella et al. 2006; Vanzella et al. 2007). There are 275 Lyman-break galaxies in both the GOODS fields which are classified as V −band "dropouts" i.e. z ∼ 5. The magnitude limit of the GOODS optical observations imply that they are brighter than ∼ 0.2 L∗,UV,z=3 (Giavalisco et al. 2004), Of these, ∼20% have spectroscopic redshifts while ∼30% are indi- -- 5 -- vidually detected with IRAC. At higher redshifts, z ∼ 6, it has been shown that galaxies which are individually undetected with IRAC appear to harbor a younger stellar population and have a factor of 10 lower stellar mass than IRAC detected galaxies (Yan et al. 2006). As shown in Figure 2, GRB host galaxies are factors of 2 − 3 times fainter than the median V −band luminosity of galaxies which have spectroscopic redshifts of 4.5 < z < 5.5 in the GOODS field. Furthermore, the luminosities are comparable to the rest-frame V −band luminosity of GRB hosts studied at lower redshifts (e.g.; Chary et al. 2002; Le Floc'h et al. 2003). This suggests that GRB host galaxies are unlike the luminous end of the star-forming, Lyman-break galaxy population which have had about a factor of 10 increase in their stellar mass between z ∼ 5 and z ∼ 1. They are more typical of the blue, faint end of the galaxy V −band luminosity function, a population for which it is difficult to measure redshifts or metallicities, in the absence of GRBs, due to their inherent faintness. GRB host galaxies at z ∼ 0.5 − 3, which have extensive multi-wavelength data, show clear evidence for very high specific star formation rates indicating an on-going starburst (Chary et al. 2002; Christensen et al. 2004; Castro Cer´on et al. 2006). We do not yet have constraints on the star formation rates in the z ∼ 5 host galaxies presented here, due to their intrinsic faintness in the rest-frame UV (see e.g. Fruchter et al. 2006; Jakobsson et al. 2005). However, spectroscopy of the afterglows by Price et al. (2007) and Berger et al. (in prep.) has revealed a wealth of absorption lines which have been used to derive the metallicity and gas column density in the vicinity of the burst. Absorption spectroscopy of the three bursts presented here have yielded neutral hydro- gen gas densities in their host galaxies of: log[N(HI)] = 21.6 ± 0.1 (060223A), log[N(HI)] = 21.3 ± 0.1 (060510B), and log[N(HI)] = 21.0 ± 0.3 (060522). Thus, all three systems are clearly DLAs, with column densities near the median of the distribution for GRB-DLAs (Berger et al. 2006b; Jakobsson et al. 2006). In addition, the metallicities of the GRB 060223A and 060510B systems have been determined from the detection of weak metal lines. For GRB 060522 the signal-to-noise of the spectrum is too low to clearly identify any metal lines and an estimate of the metallicity is thus not possible. In the case of GRB 060223A, we find an upper limit on the column density of S II of log[N(SiII)] < 15.3, leading to a metal- licity of [S/H] < −1.45. The non-detection of Fe IIλ1608 leads to a limit of [Fe/H] < −2.65, but we stress that iron can be heavily depleted onto dust grains. From the detection of the Si IIλ1304 line we find log[N(SiII)] ≈ 15.3, and hence [Si/H] ≈ −1.8. As in the case of iron, silicon is also strongly depleted, so we conclude that the metallicity of the GRB 060223A DLA is in the range of ∼ −1.8 to ∼ −1.4. For GRB 060510B, we use the S IIλλ1250, 1253 lines to measure log[N(SiII)] = 15.6 ± 0.1, and hence a metallicity, [S/H] = −0.85 ± 0.15 (see also Price et al. 2007). -- 6 -- The metallicity estimates of the GRB hosts along with their rest-frame B−band lumi- nosities (assuming a B − V color of 0, typical of star-forming galaxies) are shown in Figure 3. Also shown for comparison are the metallicity-luminosity relationships for different sam- ples of field galaxies. Despite the one detection and two limits for the luminosity of the host galaxies, the figure shows that the redshift evolution of metallicity at a fixed B−band luminosity that is seen between 0 < z < 2, clearly extends out to z ∼ 5. The chemical enrichment of galaxies is directly related to their past history of star- formation since supernovae and stellar winds are responsible for recycling the products of nucleosynthesis back into the interstellar medium. The stellar mass density is the time integral of the star-formation history. By comparing the redshift evolution of the stellar mass density (ρ∗) (e.g Dickinson et al. 2003; Yan et al. 2006; Stark et al. 2007; Chary et al. 2007), with the redshift evolution of the metallicity (Z), we can search for evolution of the stellar initial mass function and assess the role of feedback in the build-up of galaxies. The Spitzer observations of the hosts are crucial, since they enable metallicity comparisons to be made at a fixed rest-frame V −band luminosity, over a wide range of redshifts. Due to the fact that we have constraints on the V −band luminosity and metallicity of only one z ∼ 5 GRB host, we make the assumption that the median metallicity at each redshift, is that of a galaxy which has a similar luminosity as the GRB host. This is not an unreasonable assumption. Within the observational uncertainties, the slope of the mass- metallicity relation appears to be invariant between z ∼ 0 and z ∼ 2 (Erb et al. 2006). The metallicity values are obtained by effectively making a vertical cut at −20.8 mag in Figure 3 and are determined to be −0.85±0.15, −0.35±0.1 and 0.33±0.1 dex at redshifts of 5, 2.3 and 0 respectively. The average estimated ρ∗ at these redshifts are 1.4, 6 and 56 in units of 107 M⊙ Mpc−3 (See references above). We performed a Monte-Carlo analysis to obtain the best fit between Z(z) and ρ∗(z). Star-forming galaxies which fall on the local mass-metallicity relationship, show a scatter of ∼0.1 dex at bright luminosities and ∼0.2 dex at faint luminosities (Tremonti et al. 2004). We use a random number generator to offset the stellar mass density and metallicity by the observed scatter from the mean values quoted above (See Dickinson et al. 2003, Table 3 for the range in stellar mass density). We fit for the relation between Z(z) − ρ∗(z) and repeat the process 10000 times. We find that dZ/dρ∗ appears to be invariant between 0 < z < 5 and that Z(z) ∝ ρ∗(z)0.69±0.17. This suggests that the chemical enrichment of star-forming galaxies takes place at a slower rate than the build up of stellar mass. This is presumably due to the loss of metals from low-mass galaxies by outflows and stellar winds, an effect which is primarily responsible for the mass-metallicity relation seen in the local Universe (Tremonti et al. 2004) and z ∼ 2 Lyman-break galaxies (Erb et al. 2006). However, alternate -- 7 -- mechanisms such as depletion of metals onto dust grains cannot be ruled out at this time. There is the possibility of a selection effect in this analysis. If long duration GRBs arise in collapsars, they might preferentially be in low-metallicity galaxies. As a result, it is possible that GRB hosts have a lower metallicity than the average field galaxy of the same rest-frame optical luminosity. Although GRB hosts appear to be have low luminosities in the rest-frame UV and V −band, the observational evidence does not indicate that the hosts have an unusually low-metallicity for their luminosity. Metallicity of GRB host galaxies appear to span the range 0.1−1 Zsolar (e.g Berger et al. 2007, 2006b; Prochaska et al. 2007b) and some of the hosts have even been found to be associated with dusty, infrared luminous galaxies (e.g Le Floc'h et al. 2006). Nevertheless, we assess the reliability of our derived Z(z) − ρ∗(z) relation by considering a bias in the metallicity of GRB environments. If we assume that the metallicity of the GRB environment is higher by >0.3 dex compared to the mean metallicity of a galaxy at its luminosity, it implies that the mean metallicity at z ∼ 5 for a field galaxy at the luminosity of the GRB host is -1.15±0.15. The best fit relation to the three points is then consistent with an exponent of unity i.e. Z(z) ∝ ρ∗(z)0.85±0.19 but has a worse χ2. The corollary is that if GRB hosts were biased by 0.3 dex towards lower metallicities, compared to the mean metallicity of a galaxy at its luminosity, the best fit relation is Z(z) ∝ ρ∗(z)0.52±0.16 which is a larger deviation from unity. Furthermore, if there were a bias in GRB host metallicities, the slope of the Z(z) − ρ∗(z) relation derived above at a fixed B−band luminosity, would have a different value between 2 < z < 5 and 0 < z < 2 due to the fact that the 0 < z < 2 relation is determined from star-forming galaxies while the 2 < z < 5 relation is derived from GRB hosts and Lyman-break galaxies. This is inconsistent with our fits, although larger samples of GRB hosts are needed to eliminate suggestions of bias. Detection of individual GRB hosts at high redshifts is likely to remain difficult, due to their intrinsic faintness. There is a clear need for homogeneous infrared surveys of GRB host galaxies which will enable stacking to be performed as a function of metallicity, gas density and rest-frame ultraviolet properties. Within our sample, GRB050904 is a marginal IRAC detection (Berger et al. 2006a), while GRB060223A is dominated by detector systematics. As a result, we are unable to provide additional constraints using stacking. Observations of a larger sample of GRB hosts, such as those currently being targeted in Spitzer program GO4-40599 (PI: Chary), will allow the luminosity-metallicity relation to be measured at high redshift and lead to a better understanding of the faint end of the galaxy luminosity function, a regime which is currently inaccessible even through ultradeep surveys like GOODS and the UDF. -- 8 -- 4. Evolution of the Star-formation Rate Density It is now well known from various mid-infrared, far-infrared and submillimeter surveys, that the star-formation rate density at z ∼ 0.5 − 2.5 is dominated by infrared luminous galaxies with LIR=L(8 − 1000 µm) > 1011 L⊙ and LIR/LUV ∼ 10 − 100 (e.g Takeuchi et al. 2005; Burgarella et al. 2006; Chary & Elbaz 2001). At z & 3, current long-wavelength sur- veys, due to their limited sensitivity, are unable to detect galaxies which harbor the bulk of the star-formation. Thus, rest-frame ultraviolet observations of galaxies are the only avenue for probing star-formation at high redshifts. The primary uncertainties associated with quantifying the SFRD at z > 3, are the contribution from galaxies at the faint end of the UV-luminosity function and dust correc- tions. Since sub-L∗,UV,z=3 galaxies contribute ∼90% of the SFRD, measurement of the faint- end slope of the ultraviolet luminosity function, where completeness corrections and surface brightness dimming issues are significant, needs to be undertaken carefully (Steidel et al. 1999; Bouwens et al. 2006). Similarly, if extinction were a significant issue, the galaxies that dominate the star-formation rate density would be UV-faint or undetected in magnitude limited rest-frame ultraviolet surveys. GRBs are relatively insensitive to these limitations. If the GRB rate density were correlated with the co-moving star-formation rate density at lower redshifts, where cross-calibration between the UV and IR are in broad agreement, measurement of the GRB rate density at z > 3 could provide an independent pathway to quantifying the SFRD (see also e.g.; Price et al. 2006). The three parameters which are most likely to dominate the calibration between GRBs and the star-formation rate density are the evolution of metallicity with redshift, evolution of the initial mass function of stars and identification and spectroscopic follow up of the GRB afterglow. If long duration GRBs were to preferentially occur in low metallicity environments, the increase in the average metallicity of the Universe with decreasing redshift would result in a higher SFR/GRB-rate ratio at low redshift. Similarly, evolution of the stellar initial mass function from a "top-heavy" to a Salpeter mass function with decreasing redshift would increase the SFR/GRB-rate ratio at low redshift. On the other hand, the detection efficiency and spectroscopic completeness of GRBs should be increasing with decreasing redshift, implying a lower SFR/GRB-rate ratio at low redshift. Calibrating each of these parameters individually is challenging at the present time, partly because the relationship between GRB rate and environment is not well known and due to the fact that observational selection effects cannot be quantified. Therefore, we need to rely on empirical comparisons between known star-formation rate estimates and GRB rate densities to assess GRBs as a star-formation rate indicator. This empirical comparison can be optimally done at z < 3 since in this redshift range the star-formation rate, including -- 9 -- the dust obscured component, has been accurately determined from deep mid-infrared and submillimeter surveys. We use the star-formation rate density at z < 3 from Chary & Elbaz (2001). We distribute the 52 Swift GRBs with spectroscopic redshifts into redshift bins and divide by the co-moving volume in each redshift bin. We also correct for the time dilation to estimate the comoving GRB rate density over the ∼ 2 year Swift lifetime. The redshift bin at z < 0.5 is omitted since the GRB rate density appears to be anomalously high compared to the rapidly evolving star-formation rate density. We find that within the uncertainties, the rate density of GRBs with spectroscopic redshifts in the range 0.5 < z < 3 is constant at a value of (3.7±1.1)×10−11 Mpc−3 yr−1. This can be compared with the extinction-corrected comoving star-formation rate density in the same redshift range which is in the range 0.12−0.25 M⊙ yr−1 Mpc−3 and has an average value of ∼ 0.2 M⊙ yr−1 Mpc−3 (Chary & Elbaz 2001). Since these two independent rate densities are relatively constant in the 0.5 < z < 3 range, we can tentatively make the assumption that the SFR/GRB-rate is constant (Figure 4). The ratio of these two rates implies: SFRD = GRB rate × (5.2 ± 2.3) × 109, (1) where SFRD is the extinction-corrected star-formation rate density in M⊙ yr−1 Mpc−3 and GRB rate is in units of Mpc−3 yr−1. Using our derived calibration, and the measured GRB rate densities at 4 < z < 5 and 5 < z < 7 of (2.4 ± 1.2) × 10−11 and (1.8 ± 0.9) × 10−11 Mpc−3 yr−1, respectively, we infer a net star-formation rate, corrected for extinction, of 0.12 ± 0.06 and 0.09 ± 0.05 M⊙ yr−1 Mpc−3 at z ∼ 4.5 and z ∼ 6 respectively. These estimates are systematically higher than those derived by Price et al. (2006) by factors of 3 − 5. The Price et al. (2006) estimates were calibrated at z ∼ 3 where neither the completeness correction factor for the faint end of the UV luminosity function nor the dust extinction correction are reliably known while deep Spitzer mid-infrared surveys have confirmed the dominant contribution of infrared luminous galaxies to the star-formation rate density at 0.5 < z < 3 (Takeuchi et al. 2005; Daddi et al. 2007). The fact that the GRB rate density is almost flat between 0.5 < z < 6, while parameters such as the detection efficiency and spectroscopic completeness should be decreasing with increasing redshift, implies that the measured GRB rate density provides at least a lower limit to the star-formation rate density. It is illustrative to compare this star formation rate estimate with those from deep rest-frame ultraviolet surveys at z > 4. Giavalisco et al. (2004) derive a star-formation rate density at z ∼ 4 of 0.02 M⊙ yr−1 Mpc−3 when integrating to 0.2 L∗,UV,z=3. After application of an extinction correction of AV = 0.45 mag, based on the extinction properties in local -- 10 -- starburst galaxies, they estimate the total star-formation rate density at z ∼ 4 to be 0.15 M⊙ yr−1 Mpc−3. Similarly, Bouwens et al. (2006) derive a star-formation rate density at z ∼ 6 by integrating the luminosity function of Lyman-break galaxies in the UDF and other deep fields. Integrating the UV luminosity function down to 0.2 L∗,UV,z=3 results in a value of 1.3 × 10−2 M⊙ yr−1 Mpc−3 while the integral to 107 L⊙ yields a SFRD of 0.04 M⊙ yr−1 Mpc−3. Application of an extinction correction, inferred to be about AUV=0.45 mag at z ∼ 6, to this latter number, implies a SFRD of 0.06 M⊙ yr−1 Mpc−3. The agreement between the SFRD values estimated from ultraviolet surveys and the GRB rate density is reassuring, considering that there have been only 8 GRBs that have been spectroscopically confirmed to be at z > 4 (Figure 4). However, the SFRD from GRBs primarily traces the faint end of the galaxy luminosity function while the surveys are measuring the contribution from the bright end. As a result, a more reasonable SFRD estimate requires adding the SFRD contribution estimated from the faint end of the galaxy luminosity function, from GRBs, to that from bright LBGs. GRB hosts are fainter than 0.2 L∗,V,z=3. Based on the UV to V −band flux ratios of star-forming galaxies at z ∼ 3, it implies that GRB hosts must be fainter than 0.2 L∗,UV,z=3 Adding the SFRD from L > 0.2 L∗,UV,z=3 galaxies to that inferred from the GRB rate density results in an extinction corrected SFRD of 0.27 ± 0.13 and 0.11 ± 0.05 M⊙ yr−1 Mpc−3 at z ∼ 4.5 and z ∼ 6, respectively. If confirmed through a larger statistical sample, this is a substantial upward revision suggesting that L < 0.2 L∗,UV,z=3 galaxies contribute at least four times as much to the star-formation rate density at z ∼ 6 as the bright end (L > 0.2 L∗,UV,z=3) of the UV luminosity function. Indirectly, this implies that the faint end slope of the UV luminosity function at z ∼ 6 must be ∼ −1.9, compared to the value of −1.73 that was derived by Bouwens et al. (2006). GRBs are a powerful tool for measuring the high redshift star-formation rate density. In particular, deep Spitzer observations of GRB hosts can reveal the contribution to the star-formation rate density from the faint end of the galaxy luminosity function, a regime which is inaccessible to deep, rest-frame ultraviolet/near-infrared surveys. Increasing the sample of high redshift GRBs will reduce the uncertainties in the star-formation rate density unaffected by extinction and through stacking analysis on the host galaxies will help estimate the contribution to the stellar mass density from sub-L∗ galaxies. Comparison between star- formation rate estimates from GRBs with those from deep UV surveys will provide better constraints on the evolution of dust extinction at high redshift and provide tremendous insights into the chemical enrichment of the early Universe. We wish to thank Mark Dickinson for his comments which strengthened the arguments -- 11 -- in this paper. We also acknowledge the extensive resources that are invested by the entire GRB community, not all of whom can be cited here, which enables prompt imaging and spectroscopic follow-up of the bursts. This work is based on observations made with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. Support for this work was provided by NASA through an award issued by JPL/Caltech. EB acknowledges support by NASA through Hubble Fellowship grant HST-01171.01 awarded by STSCI, which is operated by AURA, Inc., for NASA under contract NAS5-26555. REFERENCES Berger, E., Chary, R., Cowie, L. L., Price, P. A., Schmidt, B. P., Fox, D. B., Cenko, S. B., Djorgovski, S. G., Soderberg, A. M., Kulkarni, S. R., McCarthy, P. J., Gladders, M. D., Peterson, B. A., & Barger, A. J. 2006a, ArXiv Astrophysics e-prints Berger, E., Fox, D. B., Kulkarni, S. R., Frail, D. A., & Djorgovski, S. G. 2007, ApJ, 660, 504 Berger, E., Penprase, B. E., Cenko, S. B., Kulkarni, S. R., Fox, D. B., Steidel, C. C., & Reddy, N. A. 2006b, ApJ, 642, 979 Bouwens, R. J., Illingworth, G. D., Blakeslee, J. P., & Franx, M. 2006, ApJ, 653, 53 Bunker, A. J., Stanway, E. R., Ellis, R. S., & McMahon, R. G. 2004, MNRAS, 355, 374 Burgarella, D., P´erez-Gonz´alez, P. G., Tyler, K. D., Rieke, G. H., Buat, V., Takeuchi, T. T., Lauger, S., Arnouts, S., Ilbert, O., Barlow, T. A., Bianchi, L., Lee, Y.-W., Madore, B. F., Malina, R. F., Szalay, A. S., & Yi, S. K. 2006, A&A, 450, 69 Castro, S., Galama, T. J., Harrison, F. A., Holtzman, J. A., Bloom, J. S., Djorgovski, S. G., & Kulkarni, S. R. 2003, ApJ, 586, 128 Castro Cer´on, J. M., Micha lowski, M. J., Hjorth, J., Watson, D., Fynbo, J. P. U., & Goros- abel, J. 2006, ApJ, 653, L85 Chary, R., Becklin, E. E., & Armus, L. 2002, ApJ, 566, 229 Chary, R. & Elbaz, D. 2001, ApJ, 556, 562 Chary, R. et al. 2007, ApJ, submitted Chary, R.-R., Stern, D., & Eisenhardt, P. 2005, ApJ, 635, L5 -- 12 -- Chen, H.-W., Prochaska, J. X., Bloom, J. S., & Thompson, I. B. 2005, ApJ, 634, L25 Christensen, L., Hjorth, J., & Gorosabel, J. 2004, A&A, 425, 913 Daddi, E., Dickinson, M., Morrison, G., Chary, R., Cimatti, A., Elbaz, D., Frayer, D., Renzini, A., Pope, A., Alexander, D. M., Bauer, F. E., Giavalisco, M., Huynh, M., Kurk, J., & Mignoli, M. 2007, ArXiv e-prints, 705 Dickinson, M., Papovich, C., Ferguson, H. C., & Budav´ari, T. 2003, ApJ, 587, 25 Erb, D. K., Shapley, A. E., Pettini, M., Steidel, C. C., Reddy, N. A., & Adelberger, K. L. 2006, ApJ, 644, 813 Fazio, G. G. et al. 2004, ApJS, 154, 10 Fruchter, A. S., Levan, A. J., Strolger, L., Vreeswijk, P. M., Thorsett, S. E., Bersier, D., Burud, I., Castro Cer´on, J. M., Castro-Tirado, A. J., Conselice, C., Dahlen, T., Ferguson, H. C., Fynbo, J. P. U., Garnavich, P. M., Gibbons, R. A., Gorosabel, J., Gull, T. R., Hjorth, J., Holland, S. T., Kouveliotou, C., Levay, Z., Livio, M., Metzger, M. R., Nugent, P. E., Petro, L., Pian, E., Rhoads, J. E., Riess, A. G., Sahu, K. C., Smette, A., Tanvir, N. R., Wijers, R. A. M. J., & Woosley, S. E. 2006, Nature, 441, 463 Fynbo, J. P. U., Starling, R. L. C., Ledoux, C., Wiersema, K., Thone, C. C., Sollerman, J., Jakobsson, P., Hjorth, J., Watson, D., Vreeswijk, P. M., Møller, P., Rol, E., Gorosabel, J., Naranen, J., Wijers, R. A. M. J., Bjornsson, G., Castro Cer´on, J. M., Curran, P., Hartmann, D. H., Holland, S. T., Jensen, B. L., Levan, A. J., Limousin, M., Kouveliotou, C., Nelemans, G., Pedersen, K., Priddey, R. S., & Tanvir, N. R. 2006, A&A, 451, L47 Gehrels, N. et al. 2004, ApJ, 611, 1005 Giavalisco, M., Dickinson, M., Ferguson, H. C., Ravindranath, S., Kretchmer, C., Moustakas, L. A., Madau, P., Fall, S. M., Gardner, J. P., Livio, M., Papovich, C., Renzini, A., Spinrad, H., Stern, D., & Riess, A. 2004, ApJ, 600, L103 Hjorth, J., Møller, P., Gorosabel, J., Fynbo, J. P. U., Toft, S., Jaunsen, A. O., Kaas, A. A., Pursimo, T., Torii, K., Kato, T., Yamaoka, H., Yoshida, A., Thomsen, B., Andersen, M. I., Burud, I., Castro Cer´on, J. M., Castro-Tirado, A. J., Fruchter, A. S., Kaper, L., Kouveliotou, C., Masetti, N., Palazzi, E., Pedersen, H., Pian, E., Rhoads, J., Rol, E., Tanvir, N. R., Vreeswijk, P. M., Wijers, R. A. M. J., & van den Heuvel, E. P. J. 2003, ApJ, 597, 699 -- 13 -- Hu, E. M. & Cowie, L. L. 2006, Nature, 440, 1145 Hu, E. M., Cowie, L. L., Capak, P., McMahon, R. G., Hayashino, T., & Komiyama, Y. 2004, AJ, 127, 563 Jakobsson, P., Bjornsson, G., Fynbo, J. P. U., J´ohannesson, G., Hjorth, J., Thomsen, B., Møller, P., Watson, D., Jensen, B. L., Ostlin, G., Gorosabel, J., & Gudmundsson, E. H. 2005, MNRAS, 362, 245 Jakobsson, P., Fynbo, J. P. U., Ledoux, C., Vreeswijk, P., Kann, D. A., Hjorth, J., Priddey, R. S., Tanvir, N. R., Reichart, D., Gorosabel, J., Klose, S., Watson, D., Sollerman, J., Fruchter, A. S., de Ugarte Postigo, A., Wiersema, K., Bjornsson, G., Chapman, R., Thone, C. C., Pedersen, K., & Jensen, B. L. 2006, A&A, 460, L13 Jensen, B. L., Fynbo, J. U., Gorosabel, J., Hjorth, J., Holland, S., Moller, P., Thomsen, B., Bjornsson, G., Pedersen, H., Burud, I., Henden, A., Tanvir, N. R., Davis, C. J., Vreeswijk, P., Rol, E., Hurley, K., Cline, T., Trombka, J., McClanahan, T., Starr, R., Goldsten, J., Castro-Tirado, A. J., Greiner, J., Bailer-Jones, C. A. L., Kummel, M., & Mundt, R. 2001, A&A, 370, 909 Kawai, N., Kosugi, G., Aoki, K., Yamada, T., Totani, T., Ohta, K., Iye, M., Hattori, T., Aoki, W., Furusawa, H., Hurley, K., Kawabata, K. S., Kobayashi, N., Komiyama, Y., Mizumoto, Y., Nomoto, K., Noumaru, J., Ogasawara, R., Sato, R., Sekiguchi, K., Shirasaki, Y., Suzuki, M., Takata, T., Tamagawa, T., Terada, H., Watanabe, J., Yatsu, Y., & Yoshida, A. 2006, Nature, 440, 184 Kobulnicky, H. A. & Kewley, L. J. 2004, ApJ, 617, 240 Le Floc'h, E., Charmandaris, V., Forrest, W. J., Mirabel, I. F., Armus, L., & Devost, D. 2006, ApJ, 642, 636 Le Floc'h, E., Duc, P.-A., Mirabel, I. F., Sanders, D. B., Bosch, G., Diaz, R. J., Donzelli, C. J., Rodrigues, I., Courvoisier, T. J.-L., Greiner, J., Mereghetti, S., Melnick, J., Maza, J., & Minniti, D. 2003, A&A, 400, 499 Maiolino, R., Schneider, R., Oliva, E., Bianchi, S., Ferrara, A., Mannucci, F., Pedani, M., & Roca Sogorb, M. 2004, Nature, 431, 533 Makovoz, D. & Khan, I. 2005, in Astronomical Data Analysis Software and Systems XIV, ed. P. Shopbell, M. Britton, & R. Ebert, Vol. 347, 81 -- + Marchesini, D., van Dokkum, P., Quadri, R., Rudnick, G., Franx, M., Lira, P., Wuyts, S., Gawiser, E., Christlein, D., & Toft, S. 2007, ApJ, 656, 42 -- 14 -- Nagao, T., Murayama, T., Maiolino, R., Marconi, A., Kashikawa, N., Ajiki, M., Hattori, T., Ly, C., Malkan, M. A., Motohara, K., Ohta, K., Sasaki, S. S., Shioya, Y., & Taniguchi, Y. 2007, ArXiv Astrophysics e-prints Price, P. A., Cowie, L. L., Minezaki, T., Schmidt, B. P., Songaila, A., & Yoshii, Y. 2006, ApJ, 645, 851 Price, P. A., Songaila, A., Cowie, L. L., Bell Burnell, J., Berger, E., Cucchiara, A., Fox, D. B., Hook, I., Kulkarni, S. R., Penprase, B., Roth, K. C., & Schmidt, B. 2007, ApJ, 663, L57 Prochaska, J. X., Chen, H.-W., & Bloom, J. S. 2006, ApJ, 648, 95 Prochaska, J. X., Chen, H.-W., Bloom, J. S., Dessauges-Zavadsky, M., O'Meara, J. M., Foley, R. J., Bernstein, R., Burles, S., Dupree, A. K., Falco, E., & Thompson, I. B. 2007a, ApJS, 168, 231 Prochaska, J. X., Chen, H.-W., Dessauges-Zavadsky, M., & Bloom, J. S. 2007b, ArXiv Astrophysics e-prints Prochaska, J. X., Gawiser, E., Wolfe, A. M., Castro, S., & Djorgovski, S. G. 2003, ApJ, 595, L9 Savaglio, S., Glazebrook, K., Le Borgne, D., Juneau, S., Abraham, R. G., Chen, H.-W., Crampton, D., McCarthy, P. J., Carlberg, R. G., Marzke, R. O., Roth, K., Jørgensen, I., & Murowinski, R. 2005, ApJ, 635, 260 Shapley, A. E., Coil, A. L., Ma, C.-P., & Bundy, K. 2005, ApJ, 635, 1006 Shapley, A. E., Steidel, C. C., Adelberger, K. L., Dickinson, M., Giavalisco, M., & Pettini, M. 2001, ApJ, 562, 95 Songaila, A. & Cowie, L. L. 2002, AJ, 123, 2183 Stark, D. P., Bunker, A. J., Ellis, R. S., Eyles, L. P., & Lacy, M. 2007, ApJ, 659, 84 Steidel, C. C., Adelberger, K. L., Giavalisco, M., Dickinson, M., & Pettini, M. 1999, ApJ, 519, 1 Takeuchi, T. T., Buat, V., & Burgarella, D. 2005, A&A, 440, L17 Tremonti, C. A., Heckman, T. M., Kauffmann, G., Brinchmann, J., Charlot, S., White, S. D. M., Seibert, M., Peng, E. W., Schlegel, D. J., Uomoto, A., Fukugita, M., & Brinkmann, J. 2004, ApJ, 613, 898 -- 15 -- Vanzella, E., Cristiani, S., Dickinson, M., Kuntschner, H., Nonino, M., Rettura, A., Rosati, P., Vernet, J., Cesarsky, C., Ferguson, H. C., Fosbury, R. A. E., Giavalisco, M., Grazian, A., Haase, J., Moustakas, L. A., Popesso, P., Renzini, A., Stern, D., & The Goods Team. 2006, A&A, 454, 423 Vanzella, E. et al. 2005, A&A, 434, 53 -- . 2007, A&A Vreeswijk, P. M., Ellison, S. L., Ledoux, C., Wijers, R. A. M. J., Fynbo, J. P. U., Møller, P., Henden, A., Hjorth, J., Masi, G., Rol, E., Jensen, B. L., Tanvir, N., Levan, A., Castro Cer´on, J. M., Gorosabel, J., Castro-Tirado, A. J., Fruchter, A. S., Kouveliotou, C., Burud, I., Rhoads, J., Masetti, N., Palazzi, E., Pian, E., Pedersen, H., Kaper, L., Gilmore, A., Kilmartin, P., Buckle, J. V., Seigar, M. S., Hartmann, D. H., Lindsay, K., & van den Heuvel, E. P. J. 2004, A&A, 419, 927 Yan, H., Dickinson, M., Giavalisco, M., Stern, D., Eisenhardt, P. R. M., & Ferguson, H. C. 2006, ApJ, 651, 24 Yoshida, M. et al. 2006, ApJ, 653, 988 This preprint was prepared with the AAS LATEX macros v5.2. -- 16 -- Fig. 1. -- Spitzer image of the host galaxy of GRB 060510B at 15h56m29.607s,+78◦34′12.42′′ (J2000). Image is 12′′ on a side, North is up, East to the left. The left panel shows the processed mosaic while the right panel shows the image with the foreground galaxy 3.1′′ to the East subtracted. -- 17 -- Fig. 2. -- (Left) Histogram showing the distribution of observed 3.6 µm magnitudes for galaxies in the GOODS fields with spectroscopic redshifts 4.5 < z < 5.5. The solid symbols show the brightness of the GRB host galaxies observed in this paper relative to the field galaxies. (Right) GRB hosts have rest-frame V −band luminosities which are a factor of ∼ 2− 3 fainter than field galaxies at similar redshifts and provide a complementary way to study the faint end luminosity function of star-forming galaxies. Also shown as the shaded histogram is the V −band luminosities of GRB hosts at a median redshift of ∼1 (Chary et al. 2002; Le Floc'h et al. 2003) which indicate that GRB hosts span similar V −band luminosities, regardless of redshift. -- 18 -- 060510B 050904 060223A 0.5 0.4 0.3 0.2 0.1 0 −0.1 −0.2 −0.3 −0.4 −0.5 −0.6 −0.7 −0.8 −0.9 −1 −1.1 −1.2 −1.3 −1.4 −1.5 −1.6 −1.7 −1.8 ) ] H S / [ , ] / H O [ ( y t i c i l l t a e M −16.5 −17 −17.5 −18 −18.5 −19 −19.5 −20 −20.5 −21 −21.5 −22 −22.5 M(B) [AB mag] Fig. 3. -- Luminosity-metallicity relationship for star-forming galaxies at z < 2 compared with the host galaxies of z > 4 gamma-ray bursts. The galaxy data are from GDDS and CFRS at z ∼ 0.4 − 1 (circles; Savaglio et al. 2005), TKRS at z ∼ 0.3 − 1 (diamonds; Kobulnicky & Kewley 2004), DEEP2 at z ∼ 1 − 1.5 (squares; Shapley et al. 2005) and LBGs at z ∼ 2.3 (error bars; Erb et al. 2006). The gray lines represent the relation derived for z ∼ 0.1 galaxies from the Sloan Digitized Sky Survey (Tremonti et al. 2004). GRBs provide a unique window into the evolution of the mass-metallicity relation at high redshift and indicate that the chemical enrichment of galaxies with redshift occurs at a lower rate than the build up of stellar mass, presumably due to the expulsion of metals in low-mass galaxies by outflows. -- 19 -- Fig. 4. -- Star-formation rate density inferred from spectroscopically confirmed long-duration Swift GRBs (solid squares). The solid black line is the extinction-corrected star-formation rate density inferred at z < 4 from a variety of multiwavelength surveys in the mid- infrared and submillimeter, which are used to calibrate the GRBs (Chary & Elbaz 2001). The lower hatched region is the extinction uncorrected SFRD from rest-frame UV surveys (> 0.04 L∗,UV,z=3), including estimates by Steidel et al. (1999), Yoshida et al. (2006) and Bouwens et al. (2006). The upper hatched rectangle are these values corrected upward for reddening using the UV-slope technique by Bouwens et al. (2006). The SFRD inferred from GRBs at z > 4 is consistent, within the significant errors, to the extinction corrected SFRD. Follow-up of a larger number of high-redshift GRBs are required to confirm if the higher rate density derived from the GRBs is statistically significant. Table 1. Spitzer Observations of z ∼ 5 GRB Host Galaxies GRB RA, DEC (J2000) Redshift Date of Observation Sky Backgrounda Exposure Time GRB 060223A 03h40m49.561s,−17◦07′48.36′′ 15h56m29.607s,+78◦34′12.42′′ GRB 060510B 21h31m44.800s,+02◦53′10.35′′ GRB 060522 4.406 4.942 5.110 2006 Sep 23 2006 Oct 26 2006 Nov 23 aBackground at 24µm, dominated by the zodiacal light. MJy/sr 24.8 13.4 28.4 130×100s 130×100s 138×100s Flux Density in µJy 5.8µm 3.6µm <0.3 <2.4 0.23±0.04 <2.1 <2.4 < 0.2 -- 2 0 --
astro-ph/0603799
3
0603
2006-05-18T13:06:52
Non-Gaussianities in two-field inflation
[ "astro-ph", "hep-ph" ]
We study the bispectrum of the curvature perturbation on uniform energy density hypersurfaces in models of inflation with two scalar fields evolving simultaneously. In the case of a separable potential, it is possible to compute the curvature perturbation up to second order in the perturbations, generated on large scales due to the presence of non-adiabatic perturbations, by employing the $\delta N$-formalism, in the slow-roll approximation. In this case, we provide an analytic formula for the nonlinear parameter $f_{NL}$. We apply this formula to double inflation with two massive fields, showing that it does not generate significant non-Gaussianity; the nonlinear parameter at the end of inflation is slow-roll suppressed. Finally, we develop a numerical method for generic two-field models of inflation, which allows us to go beyond the slow-roll approximation and confirms our analytic results for double inflation.
astro-ph
astro-ph
Non-Gaussianities in two-field inflation Filippo Vernizzi1 and David Wands2 1Helsinki Institute of Physics, P.O. Box 64, FIN-00014 University of Helsinki, Finland and 2Institute of Cosmology and Gravitation, University of Portsmouth, Portsmouth PO1 2EG, United Kingdom November 16, 2018 Abstract We study the bispectrum of the curvature perturbation on uniform energy density hypersurfaces in models of inflation with two scalar fields evolving simultaneously. In the case of a separable potential, it is possible to compute the curvature perturbation up to second order in the perturbations, generated on large scales due to the presence of non-adiabatic perturbations, by employing the δN -formalism, in the slow-roll approximation. In this case, we provide an analytic formula for the nonlinear parameter fNL. We apply this formula to double inflation with two massive fields, showing that it does not generate significant non-Gaussianity; the nonlinear parameter at the end of in- flation is slow-roll suppressed. Finally, we develop a numerical method for generic two-field models of inflation, which allows us to go beyond the slow-roll approximation and confirms our analytic results for dou- ble inflation. 6 0 0 2 y a M 8 1 3 v 9 9 7 3 0 6 0 / h p - o r t s a : v i X r a 1 1 Introduction A key prediction of a period of inflation in the very early universe, is the generation of a spectrum of primordial perturbations. Such perturbations naturally arise from the zero point vacuum fluctuations in quantum fields, which are stretched to arbitrarily large scales during inflation. The distribu- tion of the primordial density perturbations thus provides an important test of any inflation model. In particular, slow-roll models of inflation generically predict an almost scale-invariant, almost Gaussian distribution of primordial density perturbations [1, 2]. Typically one calculates the power spectrum of density perturbations and the relative amplitude of gravitational waves. However, there is limited in- formation in the power spectrum over the restricted range of scales where the primordial power spectrum can be reliably inferred from observations. As a result, there has been growing interest in calculating the distribution of primordial perturbations from inflation, considering not only the power spectrum but also the bispectrum and other measures of possible deviations from purely Gaussian distribution as a possible discriminant between differ- ent models [3, 4]. For instance, it is known that the size of the bispectrum during single-field, slow-roll inflation is related to the spectral tilt of the power spectrum and is thus constrained to be small [5]. On the other hand, inflationary models with higher derivative operators, such as ghost infla- tion or inflation based on the Dirac-Born-Infeld action, can produce higher non-Gaussianity [6, 7, 8, 9], possibly detectable. Furthermore, non-adiabatic perturbations produced during multi-field inflation can certainly generate de- tectable non-Gaussianity in the density field after inflation, in models such as the curvaton scenario [10] (see also [11]) or reheating [12, 13, 14, 15]. What is less clear is whether nonlinear large-scale evolution of pertur- bations during inflation is capable of producing significant non-Gaussianity. Such a study requires a consistent treatment of nonlinear perturbations, not just in the matter fields but also in the gravitational field. Recently a num- ber of authors have developed gauge-invariant descriptions of the nonlinear curvature perturbation on large scales [16, 17, 18, 19, 20, 21]. Rigopoulos et al. [22] have used numerical simulations of the stochastic dynamics on large scales and found significant non-Gaussianity even in simple two-field models. It is this question which we wish to address in this paper by presenting analytical and numerical estimates of the non-Gaussianity in two-field inflation models. We use the δN-formalism, as advocated by Lyth and Rodriguez [20], to calculate the evolution of the curvature perturbation after Hubble exit. We find that in the case of a separable potential, it is possible to construct an analytical formula to express the bispectrum of the 2 curvature perturbation in terms of the potentials and slow-roll parameters of the two fields. Furthermore, we also use numerical solutions to go beyond the slow-roll approximation after Hubble exit, confirming our analytic results. In the case of double inflation with two massive fields, we find no significant non-Gaussianity produced, which confirms a previous discussion by Alabidi and Lyth in Ref. [23]. Note that the multi-field scenario considered here is different from that considered in [11, 24], where only one field is contributing to the energy density during inflation. In Sec. 2 we review the two- and three-point statistics of field perturba- tions during slow-roll inflation, and how in the δN-formalism these can be related to the two- and three-point statistics of the curvature perturbation on uniform density hypersurfaces in arbitrary multi-field models. In Sec. 3 we show how one can actually calculate the evolution of the primordial power spectrum and bispectrum using the slow-roll approximation in simple two- field models with a separable potential, during inflation. In Sec. 4 we discuss the issue of how the end of inflation can affect the curvature perturbation and the non-Gaussianity. In Sec. 5 we present both an analytical and numerical study of non-Gaussianity in the double quadratic inflation model studied by Rigopoulos et al. [22], going beyond the slow-roll approximation. We present our conclusions in Sec. 6. 2 Non-Gaussian perturbations in multi-field inflation Here we review some of the results concerning perturbations and non-Gaus- sianities in multi-field inflationary models. For simplicity, we will consider a model of inflation driven by a set of minimally coupled scalar fields with canonical kinetic terms and potential W , described by the action m2 P 2 Z d4x√−g"1 S = − P ≡ 8πG is the reduced squared Planck mass. where m2 ∂µϕI∂µϕI + W (ϕ1, ϕ2, . . .)# , 2 XI (1) We consider scalar perturbations in a quasi-homogeneous flat Friedmann- Lemaıtre-Robertson-Walker spacetime with scale factor a(t) and perturbed metric ds2 = −(1 + 2A)dt2 + 2B,idtdxi + a(t)2 [(1 − 2ψ)δij + 2E,ij] dxidxj. Primordial cosmological perturbations are usually expressed in terms of the curvature perturbation on uniform energy density hypersurfaces, denoted (2) 3 by ζ, defined in [25, 26] in linear perturbation theory, and generalized at higher order in the perturbations in [16, 17, 18]. (See also [19, 27, 28] for comparison between different variables at second order.) This quantity is widely used, especially because it is conserved, on large scales, for adiabatic perturbations [29]. Here we want to compute the three-point correlation function of ζ and hence we need to define ζ only up to second order in the perturbations, ζ ≡ −ψ − ψ2 − H ρ δρ + 1 ρ ψδρ + H ρ2 δρ δρ + 1 ρ !. 2 ρ H δρ2, (3) where ρ is the background energy density, δρ its perturbation, and H ≡ a/a is the Hubble rate, while a dot denotes the derivative with respect to cosmic time t. According to the so called δN-formalism [30, 31, 32], ζ, evaluated at some time t = tc, is equivalent, on large scales, to the perturbation of the number of e-foldings N (tc, t∗, x) from an initial flat hypersurface at t = t∗, to a final comoving -- or, equivalently, uniform density -- hypersurface at t = tc. Hence, one has, on large scales, ζ(tc, x) ≃ δN(tc, t∗, x) ≡ N (tc, t∗, x) − N(tc, t∗), where N(tc, t∗) ≡ Z c ∗ Hdt (4) (5) is the unperturbed value of N . Equation (4) simply follows from the def- inition of N as the volume expansion rate of the t = const hypersurface, integrated along the integral curve of the unit vector orthogonal to the t = const hypersurface [31]. This definition is not restricted to linear theory but holds also at second or higher order in the cosmological perturbations [32, 33, 18, 21]. One can then take t∗ as the time, during inflation, when the relevant perturbation scales exited the Hubble radius, k = aH, and tc as some time > t∗ during or after inflation. Then the number of e-foldings N can be viewed as a function of the field configuration ϕI(t∗, x) on the flat hypersurface at t = t∗ and of the time tc. If one splits the scalar fields in a background value and a perturbation, ϕI(t, x) ≡ ϕI(t) + δϕI(t, x), δN(tc, t∗, x) can be expanded in series of the initial field perturbations δϕI(t∗, x). Retaining only terms up to second order, one obtains δN(tc, t∗, x) = XI N,IδϕI ∗ + 1 2 XIJ N,IJ δϕI ∗δϕJ ∗ , (6) 4 where N,I ≡ ∂N ∂ϕI ∗ , N,IJ ≡ ∂2N ∗∂ϕJ ∗ ∂ϕI , (7) are the first and second derivatives of the unperturbed number of e-foldings N(tc, t∗), with respect to the unperturbed values of the fields at Hubble crossing. Note that in general N depends on the fields, ϕI(t), and their first time derivatives, ϕI(t). However, if slow-roll conditions 3H ϕI ≃ −W,I, at t = t∗ (8) are satisfied at Hubble exit, then N depends only on the field values [31]. In [20] it was shown that Eq. (4) with (6) can be used to compute ζ up to second order in the perturbations in multi-field models of inflation. In particular, the δN-formalism is equivalent to integrate the evolution of ζ on super-Hubble scales from Hubble exit until tc. 2.1 Two-point statistics The power spectrum of the curvature perturbation ζ, Pζ, is defined as hζk1ζk2i ≡ (2π)3δ(3)(k1 + k2) 2π2 1 Pζ(k1). k3 (9) Let us now take the scalar field perturbations δϕI as uncorrelated stochas- tic variables at early times, with scale invariant spectrum of massless scalar fields in de-Sitter space. Their two-point correlation functions hence satisfy hδϕI k1δϕJ k2i = (2π)3δIJ δ(3)(k1 + k2) 2π2 1 P∗(k1), k3 P∗(k) ≡ H 2 ∗ 4π2 , (10) where H∗ is evaluated at Hubble exit, k = aH. From Eqs. (9) and (10), making use of Eqs. (4) and (6), one obtains Pζ = XI N 2 ,IP∗. (11) Another interesting observable that can be derived from the two-point statistics is the scalar spectral index of Pζ, defined as nζ − 1 ≡ d lnPζ d ln k . (12) For multi-field inflation, the expression for the scalar spectral index has been given, for instance, in [31, 1]. At lowest order in slow-roll one has [2] d lnPζ d ln k ≃ d lnPζ dN = 1 H d lnPζ dt . (13) 5 The first equality follows from taking k at Hubble crossing, k = aH ∝ H exp N, while the second is a consequence of the definition of N, Eq. (5). Making use of the expression for the scalar power spectrum Pζ given in (11), and of (14) (15) (16) d dt ϕI = XI ∂ ∂ϕI , one thus obtains nζ − 1 = −2ǫ + 2 H PIJ ϕJ N,JIN,I ,K , PK N 2 where ǫ is the well known slow-roll parameter defined as H H 2 . ǫ ≡ − This expression can be shown to be equivalent to that given in [31]. Indeed, the particular combination of second derivatives, ϕJ N,IJ, that appears in the spectral index (15) can be eliminated in favor of second derivatives of the potential using the slow-roll approximation to give [1] nζ − 1 = −2ǫ − 2 m2 PPK N 2 ,K + 2m2 PPIJ V,JIN,IN,J V PK N 2 ,K . (17) 2.2 Three-point statistics Let us now discuss the three-point statistics of the curvature perturbations. The bispectrum of the curvature perturbation ζ is defined as hζk1ζk2ζk3i ≡ (2π)3δ(3)(Xi ki)Bζ(k1, k2, k3). (18) Observational limits are usually put on the so called nonlinear parameter fNL. We use here the definition of k-dependent fNL given in [5],1 6 5 − fNL ≡ Πik3 i i Pi k3 Bζ 4π4P 2 ζ . (19) One can compute the bispectrum using Eqs. (4) and (6). For the three- point correlation function of ζ, this yields k1δϕJ N,IN,J N,KhδϕI hζk1ζk2ζk3i = XIJK k2δϕK 1 2 XIJKL N,IN,JN,KLhδϕI k1δϕJ k3i + k2(δϕK ⋆ δϕL)k3i + perms, (20) 1Note that this choice of sign for fNL differs from that used by [51, 52]. The minimal detectable fNL with an ideal CMB experiment is slightly larger than unity [53, 51]. 6 where a star denotes the convolution and we have neglected correlation func- tions higher than the four-point (see [35, 34]). The first line represents the contribution from the three-point correlation functions of the fields. For purely Gaussian fields, this vanishes. However, Seery and Lidsey [35] have found that during slow-roll inflation the field three-point correlation functions do not vanish and read k1δϕJ k2δϕK hδϕI k3i = (2π)3δ(3)(Xi ki) 4π4 Πik3 i P 2 ∗ Xperms ϕIδJK 4Hm2 PM, with M = M(k1, k2, k3) ≡ −k1k2 2 − 4 k2 2k2 3 kt + 1 2 k3 1 + k2 2k2 3 k2 t (k2 − k3) , (21) (22) where kt = k1 +k2 +k3, and it is assumed that all the ki are of the same order of magnitude so that they cross the Hubble radius approximately at the same time. The function M, which has been written in a slightly different form from that in Ref. [35], parameterizes the k-dependence of the three-point functions. The sum is over all simultaneous rearrangements of the indices I, J, and K, and the momenta k1, k2, and k3 in M, such that the relative position of the ki is respected [35]. The sum over the permutations of M over all the ki reads F (k1, k2, k3) ≡ XpermsM(k1, k2, k3) = −2  1 2 Xi6=j kik2 j + 4Pi>j k2 kt i k2 j − 1 2 Xi k3 i  , (23) and one can sum over the permutations and evaluate the first line of Eq. (20) using Eq. (21) and XIJK N,IN,JN,K Xperms ϕIδJKM(k1, k2, k3) = −H XI N 2 ,IF (k1, k2, k3), (24) where we have used PI N,I ϕI = −H. To evaluate the second line of Eq. (20), as in [35] we assume that the connected part of the four-point correlation functions is negligible and we make use of Wick's theorem to reduce the four-point functions to products of two-point functions. Hence, we can finally rewrite the bispectrum using its definition (18) and Eq. (20). This yields, after few manipulations, B(k1, k2, k3) = 4π4P 2 ζ Pi k3 i i Πik3 4m2 −1 PPK N 2 ,K F Pi k3 i + PIJ N,IN,JN,IJ ,K)2 ! . (25) (PK N 2 7 To derive this expression we have used Eq. (11) to replace P 2 Eq. (19), One can relate the bispectrum to the nonlinear parameter fNL by using ∗ with P 2 ζ . 6 5 − where fNL = P∗ 2m2 (1 + f ) + PIJ N,IN,J N,IJ ,K)2 PPζ (PK N 2 f = f (k1, k2, k3) ≡ −1 − F 2Pi k3 i , (26) (27) is a function of the shape of the momentum triangle with the range of values 0 ≤ f ≤ 5 6 [5], and we have used Eq. (11) in the first term on the right hand side of Eq. (26). The lower bound on f is obtained in the geometrical limit when one of the three ki's is much smaller than the other two, e.g., k1 ≪ k2 ≈ k3, in which case fNL becomes independent of k [5, 36], while the upper bound is obtained in the equilateral triangle configuration, k1 ≈ k2 ≈ k3. The first term on the right hand side of Eq. (26) is momentum dependent and comes from the three-point correlation functions of the fields, Eq. (21), computed by quantizing the perturbations inside the Hubble radius during inflation [35, 37]. Notice that P∗ can be related to the amplitude of gravita- tional waves. Introducing the ratio between tensor to scalar modes, this term can be written as r ≡ 8P∗ m2 PPζ , 6 5 − f (3) NL ≡ r 16 (1 + f ), (28) (29) and is constrained by observations to be small, r/16 ≪ 1. This expression simplifies the bound on this term found in Ref. [37]. The second term [20], 6 5 − f (4) NL ≡ PIJ N,IN,J N,IJ ,K)2 (PK N 2 , (30) is momentum independent and local in real space, because it is due to the evolution of nonlinearities outside the Hubble radius during inflation. The nonlinear parameter fNL can only be larger than unity if the mo- NL is large. We will devote the next section to mentum independent term f (4) compute this quantity in two-field models. 8 3 Two-field inflation with separable potential We will consider now two scalar fields ϕ1 ≡ φ and ϕ2 ≡ χ, described by the action S = − m2 P 2 Z d4x√−g(cid:20)1 2 ∂µφ∂µφ + 1 2 ∂µχ∂µχ + W (φ, χ)(cid:21) . (31) We assume that the potential of the fields is separable into the sum of two potentials each of which is dependent on only one of the two fields, W (φ, χ) = U(φ) + V (χ) . (32) 3.1 Slow-roll dynamics of background fields The Klein-Gordon equations for the background fields read φ + 3H φ + U ′ = 0, χ + 3H χ + V ′ = 0, where V ′ ≡ The unperturbed Friedmann equations read U ′ ≡ , dU dφ dV dχ . H 2 = H = − χ2 + W(cid:19) , 2 1 2 φ2 + 1 3m2 1 2m2 P (cid:18)1 P (cid:16) φ2 + χ2(cid:17) . (33) (34) (35) (36) (37) Inflation takes place when ǫ = − H/H 2 < 1. Here it is assumed that the slow-roll conditions are satisfied for both fields during inflation. In this case, the Klein-Gordon and the first Friedmann equations reduce to 3H φ ≃ −U ′, 3H χ ≃ −V ′, 3m2 PH 2 ≃ W. (38) It is hence convenient to extend the definition of slow-roll parameter ǫ for one field and define [38] m2 P W!2 2 U ′ , ǫφ ≡ m2 P W!2 2 V ′ , ǫχ ≡ with ǫ = ǫφ + ǫχ. 9 (39) (40) One can also define the two slow-roll parameters ηφ ≡ m2 P U ′′ W , ηχ ≡ m2 P V ′′ W . (41) Note that throughout this paper, for simplicity, we will assume U ′ ≥ 0 and V ′ ≥ 0 so that we may eliminate first derivatives of the potential in favor of the slow-roll parameters √ǫφ and √ǫχ. We will use the slow-roll equations (38) to write the number of e-foldings (5) during inflation as [30, 1] N = − 1 m2 P Z c ∗ U U ′ dφ − 1 m2 P Z c ∗ V V ′ dχ. (42) There is a crucial difference between single field inflation and inflation with many fields [38]. In single field inflation the slow-roll solution forms a one-dimensional phase space. This means that once the inflationary attrac- tor has been reached, there is a unique trajectory. In particular, the end of inflation takes place at a fixed value of the inflaton field which in turn corresponds to a fixed energy density. However, if two fields are present, the phase-space becomes two-dimensional and there is an infinite number of possible classical trajectories in field space. The values of the two fields at the end of inflation will in general depend on the choice of trajectory. In- terestingly, for the potential (32), under the slow-roll conditions (38), there exists a dimensionless integral of motion C, using which one can label each slow-roll classical trajectory, C ≡ −m2 PZ dχ V ′ . PZ dφ U ′ + m2 (43) The number of e-foldings N in Eq. (42) characterizes the evolution along a given trajectory, while the quantity C allows us to parameterize motion off the classical trajectory, and it will turn out to be very useful in computing the curvature perturbation during inflation. In this respect, the study presented here is very similar to the one discussed in Ref. [38], in the case of a potential that is a separable product rather than a separable sum. 3.2 Perturbed expansion during slow-roll In order to calculate ζ, given in Eq. (4), we need to calculate the perturbed expansion due to the field quantum fluctuations on an initial spatially flat slice up to a final comoving or uniform density hypersurface at tc. To calculate the power spectrum of ζ (11) we require only the first derivative of the 10 expansion with respect to the initial field values. But to calculate the scalar spectral index (15) or the three-point correlation function (20) we need to compute the perturbation in the number of e-foldings expanded up to second order in the initial field perturbations. We will thus proceed as follows. We will first compute the first derivative of N(tc, t∗) with respect to the fields, i.e., N,φ and N,χ, by differentiating N(tc, t∗) in Eq. (42) in dφ∗ and dχ∗. With these two first derivatives we will be able to compute the second derivatives N,IJ . We first note that each of the integrals in Eq. (42) depends upon both φ∗ and χ∗. For instance, the value of the integral over φ depends upon χ∗ through the dependence of the limit φc upon C(φ∗, χ∗), the integral of motion (43) labelling the classical trajectory. In other words one has dN = 1 m2 1 m2 P "(cid:18) U P "(cid:18) V U ′(cid:19)∗ − V ′(cid:19)∗ − ∂χc ∂φ∗ (cid:18) V ∂χ∗ (cid:18) V V ′(cid:19)c − V ′(cid:19)c − ∂χc ∂φc ∂φ∗ (cid:18) U ∂χ∗ (cid:18) U U ′(cid:19)c# dφ∗ U ′(cid:19)c# dχ∗. ∂φc + (44) In order to compute the derivatives of the number of e-foldings with respect the initial fields φ∗ and χ∗, one needs to compute ∂χc/∂χ∗, ∂χc/∂φ∗, etc. Since the value of φc and χc on a given classical trajectory is a function of the conserved quantity C, one has dφc = dχc = dφc ∂φ∗ dC ∂C dC ∂C dχc ∂φ∗ dφ∗ + dφ∗ + ∂C ∂χ∗ ∂C ∂χ∗ dχ∗! , dχ∗! . Making use of Eq. (43), one can easily compute ∂C ∂φ∗ = − m2 P U ′ ∗ , ∂C ∂χ∗ = m2 P V ′ ∗ . (45) (46) We now require that t = tc coincides with a ρ = const hypersurface, which during slow-roll is given by U(φc) + V (χc) = const . Differentiating this condition yields dφc dC U ′ c + dχc dC V ′ c = 0. 11 (47) (48) Furthermore, differentiating Eq. (43) with respect to the trajectory C and using Eq. (48), one finally obtains m2 P m2 P dφc dC dχc dC 2 + = −"U ′ c 1 c 1 = "V ′ 2 + V ′ c V ′ c , 1 U ′ c 2!#−1 2!#−1 . 1 U ′ c Thus, substituting Eqs. (46) and (49) in Eq. (45), one finds ∂φc ∂φ∗ = Wc W∗ ∂χc ∂φ∗ = − , ǫχ c c ǫφ ∗!1/2 ǫc ǫφ ∗!1/2 ǫc ǫχ c ǫφ ǫφ c Wc W∗ ∂φc ∂χ∗ = − Wc W∗ , ∂χc ∂χ∗ = Wc W∗ ǫχ c c ǫχ ǫc ǫφ ǫc (cid:18)ǫχ ∗!1/2 ∗(cid:19)1/2 c ǫχ ǫφ c , , (49) (50) where ǫ is the overall slow-roll parameter given in Eq. (16) and we have written the derivatives of the potentials that appear in Eqs. (46) and (49) in terms of the slow-roll parameters defined in Eq. (39). One can now evaluate the derivatives of the number of e-foldings with respect to the initial fields φ∗ and χ∗, using in Eq. (44) the expressions (50). For the first derivative this yields mP mP ∂N ∂φ∗ ∂N ∂χ∗ = = where 1 U∗ + Zc ∗ q2ǫφ 1 √2ǫχ ∗ W∗ V∗ − Zc W∗ Zc = (Vcǫφ c − Ucǫχ c )/ǫc . , , (51) (52) (53) To compute the second derivatives we differentiate Eqs. (51) and (52) with respect to φ∗ and χ∗. This yields m2 P ∂2N ∂φ2 ∗ m2 P ∂2N ∂χ2 ∗ ∂2N ∂φ∗∂χ∗ m2 P = 1 − = 1 − ηφ ∗ 2ǫφ ∗ ηχ ∗ 2ǫχ ∗ = mP W∗q2ǫφ ∗ U∗ + Zc W∗ + V∗ − Zc W∗ − ∂Zc ∂χ∗ = − mP ∗ mP W∗q2ǫφ W∗√2ǫχ mP W∗√2ǫχ ∗ ∗ ∂Zc ∂φ∗ ∂Zc ∂χ∗ ∂Zc ∂φ∗ . , , (54) (55) (56) 12 In single field inflation, perturbations on large scales are adiabatic and hence ζ remains constant during inflation [29]. In two-field models, however, the large-scale perturbations are not in general adiabatic and ζ evolves dur- ing inflation. In the previous expressions, Eqs. (51), (52) and (54 -- 56), the function Zc takes account of this evolution by expressing how the local expan- sion depends upon the local field values on the uniform density hypersurface during inflation. 3.3 Curvature perturbation and non-Gaussianity dur- ing inflation We now compute the curvature perturbation and the non-Gaussianity on a comoving or uniform energy density hypersurface during slow-roll inflation, as given in Eq. (4). From the first derivatives of N, Eqs. (51) and (52), one can derive an expression for the power spectrum of the curvature perturbation (11). It can be simplified by introducing two dimensionless variables, related to the values of the potentials and first slow-roll parameters at t∗ and tc, u ≡ U∗ + Zc W∗ , v ≡ V∗ − Zc W∗ . The power spectrum then reads Pζ = W∗ 24π2m4 P u2 ǫφ ∗ + v2 ǫχ ∗! . (57) (58) Note that this can never be less than the power spectrum derived by con- sidering only adiabatic perturbations restricted to the background trajectory (at fixed C). In multi-field inflation there is an additional contribution to the power spectrum due to non-adiabatic perturbations at Hubble exit. Thus one has Pζ ≥ P ζC , where P ζC can be obtained by taking the limit tc → t∗ in Eq. (58), since for adiabatic perturbations the curvature perturbation does not change after Hubble crossing. In this limit u → ǫφ ∗ ǫ∗ , ǫχ ∗ ǫ∗ , v → for tc → t∗, and one finds P ζC = W∗ 24π2ǫ∗m4 P . (59) (60) To compute the scalar spectral index nζ with Eq. (15) and the nonlinear parameter fNL with Eq. (26) one needs the second derivatives N,IJ , given in 13 Eqs. (54 -- 56). In the expression for the scalar spectral index the derivatives of Zc with respect to the fields contained in these equations mutually cancel. Thus, by using Eq. (15), one finds nζ − 1 = −2ǫ∗ − 4 u(cid:18)1 − ηφ 2ǫφ ∗ ∗ u(cid:19) + v(cid:16)1 − ηχ 2ǫχ ∗ ∗ u2 ǫφ ∗ + v2 ǫχ ∗ v(cid:17) . (61) Note that in the limit tc → t∗, using Eq. (59), one finds the spectral index for purely adiabatic perturbations, nζC − 1 = −6ǫ∗ + 2ησσ ∗ , where we have defined [39] ησσ ≡ (ǫφηφ + ǫχηχ)/ǫ. (62) (63) This represents the effective mass of adiabatic fluctuations tangent to the background trajectory. The momentum dependent nonlinear parameter, f (3) NL, defined in Eq. (29), takes a very simple form, derived from Eq. (60), − 6 5 NL = ǫ∗PζC f (3) Pζ (1 + f ) ≤ ǫ∗(1 + f ), (64) where f is the momentum dependence, defined in Eq. (27). In order to give an expression for f (4) NL, one needs ∂Zc/∂φ∗ and ∂Zc/∂χ∗, which can be computed by employing Eq. (50). One finds qǫφ ∗ ∂Zc ∂φ∗ = −qǫχ ∗ ∂Zc ∂χ∗ = √2 mP W∗A, where we have defined with [39] A ≡ − W 2 c W 2 ∗ c ǫχ ǫφ ǫc (cid:18)1 − c ηss c ǫc (cid:19) , ηss ≡ (ǫχηφ + ǫφηχ)/ǫ. (65) (66) (67) This represents the effective mass of isocurvature fluctuations orthogonal to the background trajectory. 14 We are thus able to give an analytic expression for f (4) NL defined in Eq. (30). One finds 6 5 − f (4) NL = 2 u2 ǫφ ∗ (cid:18)1 − ηφ 2ǫφ ∗ ∗ ∗ ǫχ 2ǫχ v(cid:17) +(cid:18) u u(cid:19) + v2 ∗ (cid:16)1 − ηχ ∗(cid:19)2 (cid:18) u2 + v2 ǫχ ǫφ ∗ ∗ ∗(cid:19)2 ∗ − v ǫχ ǫφ A . (68) This is the exact expression for the amplitude of the nonlinear parameter for the curvature perturbation ζ during slow-roll inflation in an arbitrary two-field inflation model with separable potential, and represents one of the main results of this paper. As for the power spectrum and spectral index, we can take the limit of tc → t∗, which gives the nonlinear parameter for purely adiabatic perturba- tions in the two-field case. Using Eq. (59) in Eqs. (64) and (68), yields 6 5 − fNLC = ǫ∗(1 + f ) + 2ǫ∗ − ησσ ∗ (69) which coincides with the single field case result [5, 36]. Note that, using Eqs. (62) and (28) for the adiabatic case, this can be expressed as a consis- tency relation between observables [5], 6 5 − fNLC = − 1 2 (nζC − 1) + rC 8 f , (70) and is thus constrained to be small. Although the general expression (68) for two fields, which allows for non- adiabatic perturbations, is rather involved, we can qualitatively discuss the order of magnitude we expect for f (4) NL. Since max(u, v) ≤ 1 and u + v = 1, the denominator in Eq. (68) is of order ε−2 ∗ , where we use ε to denote generic first-order slow-roll parameters, ǫ or η. The first two terms in the numerator of Eq. (68) are of order ε−1 NL is of order ε∗. Only the third term, which is of order Aε−2 ∗ , leads to a contribution that is not automatically slow-roll suppressed. ∗ , so their contribution to f (4) Although ηss may become larger than unity during inflation, the prefac- tor in front of the parenthesis in Eq. (66) can be very small or vanishing, suppressing the contribution of A to the nonlinear parameter. If this is not the case, since A does not appear in the expression for the spectral index nζ, Eq. (61), for ηss c /ǫc ≫ 1 one may have models with large fNL and quasi- scale-invariant spectral index, corresponding to large deviations from the consistency relation for purely adiabatic perturbations, Eq. (70). We leave to future work the investigation of models where A, and thus fNL, are large. 15 4 Perturbations after inflation So far we have considered only the curvature perturbation during inflation. In order to relate our calculations to observables we need to calculate ζ after inflation when the universe is radiation dominated, on a uniform energy density hypersurface for t = tc > te, where te denotes the end of inflation. In this case tc defines a uniform radiation density hypersurface. In single field inflation large-scale perturbations are adiabatic and thus the curvature perturbation on uniform density hypersurfaces remains constant both during and after inflation, independent of the detailed physics occurring at the end of inflation. In two-field models we need to take account of how the local expansion depends upon the local field values both during inflation and at the end of inflation in order to calculate ζ some time after inflation. If one of the two fields -- for instance φ -- has stabilized before the end of inflation, so that the end of inflation is dominated by a single field χ, then Zc becomes constant, and the power spectrum, the scalar spectral index and the non-linear parameters after the end of inflation are simply given by Eqs. (58), (61), (64) and (68) with Zc = Uc = const. Note that in this case A = 0 at the end of inflation and the nonlinear parameter f (4) If both fields are "active" at the end of inflation, and inflation takes place on a hypersurface q(φe, χe) = const at t = te, undefined -- i.e., not necessarily a uniform energy density, one can separate δN(tc, t∗) into the sum of two pieces, the first from t = t∗ to the end of inflation t = te, the second from t = te to a uniform density hypersurface after inflation t = tc, NL will be small. δN(tc, t∗) = δN(te, t∗) + δN(tc, te). (71) The first piece, δN(te, t∗), can be computed by a similar calculation to that of Sec. 3.2. It is possible to show that δN expanded at second order is found by replacing Zc in Eqs. (51), (52) and (54 -- 56) by Ze = Ve ∂q ∂φeqǫφ e − Ue ∂q ∂χeqǫχ e! ∂q ∂φeqǫφ e + ∂q ∂χeqǫχ e!−1 . (72) Then one needs to compute δN(tc, te). In scenarios such as the curvaton or modulated reheating scenarios, it is assumed that the end of inflation hypersurface is effectively unperturbed, δN(te, t∗) ≃ 0, and that the primordial density perturbation originates from isocurvature field perturbations, δs, during inflation, which introduce a per- turbation δN(tc, te) ∝ δs only after or at the end of inflation. An alternative possibility is to assume that inflation ends due to a sudden instability triggered by some function of the fields reaching a critical value 16 [40, 41, 42]. This is what happens in hybrid inflation [43, 44] where the false vacuum state is destabilized when the inflaton field, φ, reaches a critical value, φ = φe. If we assume instantaneous reheating of the universe one has [33] δN(tc, te, x) = N (ρ(tc, x)) − N (ρ(te, x)) dρ2 δρ2#e 2 ρ 2ρ − = −" dN = −He" δρ δρ + d2N 1 2 1 ρ ρ! δρ2 ρ2 #e dρ + ρ . (73) We assume that the energy density is conserved at the end of inflation, so that the background value of the energy density in this expression is ρ(te) = U(φe) + V (χe). Furthermore, if the equation of state becomes radiation-like, so that ρ = −4Hρ, we then have δN(tc, te, x) = 1 4 " δρ ρ − 1 2 δρ2 ρ2 #e . (74) Note that the numerical coefficient that appears in front of the brackets in this expression is dependent upon the equation of state after inflation has ended and would be 1/3(1 + w) for an equation of state P = wρ. In a hybrid-type limit, where the false vacuum dominates the self-interac- tion energy of the slowly-rolling fields, we can take the effective potential to be almost completely flat (consistent with the slow-roll approximation) so that [δρ/ρ]e is negligible and δN(tc, te) ≪ δN(te, t∗). The primordial curvature perturbation ζ is then given by the curvature of the end of inflation hypersurface, ζ(tc) = δN(te, t∗). 5 Double quadratic inflation To be more specific and give a quantitative estimate of the non-Gaussianity, it is instructive to consider the case of massive fields with potential given by Eq. (32) with [45, 46] U = 1 2 m2 φφ2, V = 1 2 m2 χχ2 . (75) These scalar fields thus have no explicit interaction, but can interact gravi- tationally during inflation. 17 5.1 Slow-roll analysis We can apply our earlier analysis to calculate the curvature perturbation during slow-roll inflation in this model. The integrals (42) and (43) yield 1 1 m2 m−2 PN(tc, t∗) = 4 (cid:16)φ2 ∗ + χ2 φ ln φ∗ P C(tc, t∗) = m−2 ∗(cid:17) − c(cid:17) , 4 (cid:16)φ2 c + χ2 χc! , χ ln χ∗ φc! − m−2 (76) (77) where, as for N, we have fixed the limits of integration in the definition of C in Eq. (43) to run from t∗ to tc. In the standard treatment, one can parameterize the slow-roll trajectories of the scalar fields in polar coordinates [45] χ = 2mP√N sin θ, φ = 2mP√N cos θ. (78) The advantage of doing so is that the angular variable θ can be related to the number of e-foldings by the expression [45, 46, 47] N = N0 (sin θ)2/(R2−1) (cos θ)2R2/(R2−1) , where we have defined the ratio between the masses of the fields as R = mχ/mφ . (79) (80) Thus, Eq. (79) implies that an inflationary model is completely specified by giving R and the values of the fields φ and χ -- or equivalently N and θ -- at some given time. Once specified an initial condition, Eq. (79) can be used to evolve the background fields or equivalently the slow-roll parameters. Then one can use Eqs. (58), (61), (64) and (68), to evaluate completely analytically the power spectrum Pζ, the scalar spectral index nζ and the nonlinear parameter fNL for double inflation. Before studying in detail these equations for a specific case, one can es- timate the amount of non-Gaussianity generally produced during and after inflation. During inflation, as explained in Sec. 3.3, fNL can be large only if, in Eq. (68), the contribution proportional to the quantity A defined in Eq. (66) is large. In double inflation this contribution can become temporar- ily large, at most of order unity, at the turn of the trajectory in field space. However, after inflation, the fields have settled to their minima, which im- plies Zc = 0 and also A = 0. In this case the dependence on the field values 18 at t = tc in Eq. (57) disappears and we can simply rewrite Eqs. (58), (61), (64) and (68) by using the potential (75) and Eq. (76) with φc = χc = 0, as Pζ = H 2 ∗ 4π2m2 P nζ − 1 = −2ǫ∗ − 1 6 − 5 fNL = N(tc, t∗), 1 N(tc, t∗) , 2N(tc, t∗) [2 + f (k1, k2, k3)] , (81) (82) (83) where N(tc, t∗) is the number of e-foldings between Hubble exit and the moment when φc = χc = 0, which roughly coincides with the end of inflation. These results coincide with the estimate of Ref. [23]. We conclude that double inflation cannot produce large non-Gaussianity -- i.e., fNL ≪ 1 -- these being suppressed by the slow-roll conditions on the fields at Hubble exit. We now turn to a more detailed analysis of Eqs. (58), (61), (64) and (68), and compare with a numerical results for the evolution on large scales without assuming slow-roll. 5.2 Numerical analysis The δN-formalism assumes that the fields are in slow-roll at Hubble exit, Eq. (8) (see [48] for a development of the δN-formalism applicable to more general situations). Furthermore, in deriving Eqs. (58), (61), (64), (68) and the background evolution relation (79), we have assumed slow-roll all along the inflationary evolution, from t∗ to tc. While still assuming that slow-roll conditions are satisfied at Hubble exit, one can go beyond the slow-roll approximation by numerically solving the full background evolution equations for the fields, Eqs. (33) and (34), together with the Friedmann equation (36), and computing the evolution of the scale factor a. One can then evaluate the expansion N = ln a as a function of the field initial values (φ∗, χ∗), up to some final time tc which may be some time during or after slow-roll inflation. Then, one can numerically calculate the first and second partial derivatives of N(tc, t∗) with respect to φ∗ and χ∗ by the finite-difference method [49], i.e., by computing N for different values close to (φ∗, χ∗). Indeed, this provides an efficient method to compute the non-Gaussianity numerically for any two-field model, also when the field potential is not separable. We study, as an example, the specific case considered by Rigopoulos et al. in [22]: mass ratio R = 1/9 and initial condition φ = χ = 13mP. We assume that inflation ends on a hypersurface ǫe = 1 and that Hubble exit 19 P m / χ 16 14 12 10 8 6 4 2 0 0 2 4 6 8 φ/mP 10 12 14 16 Figure 1: Examples of trajectories in field space are shown from Hubble exit, starting 60 e-foldings before the end of inflation at ǫe = 1, for R = 1/9. The thick (orange) line represents the trajectory starting from φ = χ = 13mP, and shown from Hubble exit, (φ∗ = 8.2, χ∗ = 12.9), to the end of inflation, (φe = 0.0, χe = 1.4). The grey shading represents the space of all possible trajectories. 20 14 12 10 8 6 4 2 0 P m / χ , P m / φ 0.2 0.4 0.6 H/mφ 0.8 1 Figure 2: The values of the fields φ (solid line) and χ (dashed red line) during the inflationary trajectory of Fig. 1, are shown as a function of the Hubble rate H, from N ≃ 42 to N ≃ 4 e-foldings from the end of inflation. Note that time increases from right to left. 2 1.5 1 0.5 2 ) φ m P m ( / χ ρ , 2 ) φ m P m ( / φ ρ 0 0.2 0.4 0.6 H/mφ 0.8 1 Figure 3: The energy densities of the fields ρφ (solid line) and ρχ (dashed red line) normalized to (mPmφ)2, are shown as for Fig. 2. 21 ) 2 π 4 / 2 ∗ Η ( / ζ P 60 50 40 30 20 0.2 0.4 0.6 H/mφ 0.8 1 Figure 4: The power spectrum Pζ of the large-scale uniform density pertur- bation ζ during the inflationary trajectory of Fig. 1, is shown as a function of the Hubble rate H, from N ≃ 42 to N ≃ 4 e-foldings from the end of inflation (mP = 1). The solid and the dashed lines represent the analytic and numerical calculations, respectively. 0.93 0.92 ζ n 0.91 0.9 0.89 0.2 0.4 0.6 H/mφ 0.8 1 Figure 5: The spectral index nζ is shown as for Fig. 4. 22 0.25 0.2 0.15 0.1 0.05 L N f ) 5 / 6 ( - 0 0.2 0.4 0.6 H/mφ 0.8 1 Figure 6: The nonlinear parameter fNL, in the limit where f (k1, k2, k3) = 0 (k1 ≪ k2 ≈ k3) is shown as for Fig. 4. The (turquoise) dotted line represents NL/(1 + f ) while the (blue) dashed-dotted line represents −(6/5)f (4) −(6/5)f (3) NL, both computed analytically. takes place at t = t∗, corresponding to 60 e-foldings before the end of infla- tion. One finds that (φe = 0.0, χe = 1.4), corresponding to θe = π/2, and that (φ∗ = 8.2, χ∗ = 12.9), corresponding to θ∗ = 0.32π. The trajectory in field space is shown in Fig. 1. One can see that assuming slow-roll from t∗ to te is well justified. Indeed, the turn of trajectory in field space, corresponding to the moment when φ exits slow-roll, takes place near φ = 0, when the energy density of φ has already become negligible. This can also be checked in Figs. 2 and 3, where the values of the fields and their energy densities are shown, respectively, at the turn of the trajectory in field space, from N ≃ 42 to N ≃ 4 e-foldings from the end of inflation. NL analytically, using Eqs. (58), (61), (64) and (68), and numerically, using the procedure explained above. In Figs. 4 -- 6 we compare the analytic and numerical calculation of the power spectrum Pζ, the scalar spectral index nζ, the nonlinear parameter fNL, as a function of the Hubble rate during inflation. The two calculations agree remarkably well, within an accuracy that depends on the numerical precision. As expected, when the heavy field φ dominates the total energy density, Now we compute Pζ, nζ and fNL = f (3) NL +f (4) for H/mφ >∼ 0.8, the uniform density curvature perturbation ζ remains con- stant. It starts growing when H drops to 0.55 <∼ H/mφ <∼ 0.8, corresponding to both the light and heavy fields contributing to the total energy density. 23 During this phase, the large scale total entropy perturbation is non-vanishing, and sources the evolution of ζ [38]. After few oscillations, the heavy field en- ergy density is red-shifted and the light field dominates the universe while ζ becomes constant in time again. Consequently, the power spectrum and the spectral index grow monotonically, following the growth of ζ. The evolution of the nonlinear parameter −f (4) NL, however, is non-mono- tonic. It grows sharply during the intermediate phase 0.55 <∼ H/mφ <∼ 0.8, corresponding to the heavy field leaving slow-roll, but it then decreases. This growth corresponds to the growth of A defined in Eq. (66). One can compute the value of fNL at the end of inflation, for ǫe ≃ 1, in the limit where f (k1, k2, k3) = 0 (k1 ≪ k2 ≈ k3). One finds −(6/5)f (4) NL = 0.008 and −(6/5)f (3) NL = 0.008 which coincide with the analytical result of Eq. (83), −(6/5)fNL = 0.016 ≃ 1/60. We conclude that fNL in this model is much smaller that unity. Note that, as shown in Fig. 4, the curvature perturbation becomes con- stant after the energy density of the more massive field (φ in our case) be- comes negligible. The isocurvature perturbations in the massive field are suppressed at the end of inflation and hence the nonlinear parameter that we have calculated is indeed the primordial fNL constrained by observations, unless we have some curvaton-type mechanism which alters the large-scale curvature perturbation after inflation. 6 Conclusion In this work we have studied the non-Gaussianity, in terms of the bispectrum, generated by models of inflation with two fields evolving during inflation. We have first reviewed the use of the δN-formalism to compute the curvature perturbation on uniform density hypersurfaces during inflation, up to second order in the perturbations, when the slow-roll conditions are satisfied by all fields. We have then specialized to the two-field case, and assumed that the potential is separable into the sum of two potentials, each of which is depen- dent on only one of the two fields. In this case the number of e-foldings can be analytically expanded up to second order in the initial field fluctuations and we have derived an analytic formula, Eq. (68), to compute the nonlinear parameter fNL in terms of the values of the potential and slow-roll param- eters of the fields. Using this formula, one can identify the conditions for the model to generate large non-Gaussianity. In particular, we have shown that a large fNL after inflation requires a large ηss -- defined in Eq. (67) -- at the end of inflation, although a large ηss does not necessarily imply large 24 non-Gaussianity. As a specific model, we have considered double inflation with two massive scalar fields. In this case, the background evolution of the fields can be computed analytically and we have been able to compute the evolution of the nonlinear parameter during inflation. As shown in Sec. 5.1, one expects the non-Gaussianity generated in double inflation to be slow-roll suppressed, fNL ≪ 1, as in the single field case. Since our analytic formula relies on the slow-roll conditions of all fields from Hubble exit until the end of inflation, we have extended our analysis by developing a numerical method, based on the δN-formalism, to compute the curvature perturbation up to second order in the perturbations, that allows us to relax the slow-roll assumption after Hubble exit. This method, which can be extended to any two-field model, confirms our analytic results. Our results agree with a previous discussion of non-Gaussianity in a dou- ble quadratic potential by Alabidi and Lyth [23], based on an estimate of δN given in [1] using the integral (42). In Ref. [1] it is argued that N in Eq. (42) is dominated by the field values at Hubble exit and hence one can neglect the dependence of N upon the field values at the final time, tc. This is not in general true when evaluating the perturbation ζ during inflation. Including the dependence of φc and χc upon the field values at Hubble exit considerably complicates our analysis, but allows to follow perturbations during inflation. In our numerical example we see that the non-linearity parameter does evolve during inflation, especially as the trajectory turns the corner in field space, reflecting its dependence upon the final field values. However this does not alter the general conclusion that the non-linearity parameter remains small at the end of inflation. Our results disagree with those of Rigopoulos et al. [22] who found that double quadratic inflation, in particular the model studied in Sec. 5.2, can generate large non-Gaussianity. The formalism used by Rigopoulos et al. dif- fers from the δN-formalism in two aspects: First, it takes into account the sub-Hubble evolution of the field perturbations via stochastic terms. Second, it integrates explicitly the coupled evolution of the adiabatic and entropy super-Hubble perturbations. However, we have explicitly calculated in this model the non-Gaussianity due to the three-point function for the field per- turbations at Hubble exit, using the result of Seery and Lidsey [8], and shown that it is small, as previously argued by Lyth and Zaballa [37]. Furthermore, the δN-formalism, on super-Hubble scales, is equivalent to integrating the evolution of ζ. Thus, the contribution to fNL coming from the super-Hubble evolution should be equally taken into account by both formalisms and we are unable to explain the discrepancy between the results. We conclude, on the basis of our analysis, that nonlinear evolution on 25 super-Hubble scales during double quadratic inflation does not appear to be capable of generating a detectable level of non-Gaussianity in the bispec- trum. However, the possibility of producing a large non-Gaussianity is left open in models of two-field inflation where the mass of the isocurvature field orthogonal to the classical trajectory becomes larger than the Hubble rate and both the first slow-roll parameters of the two fields do not vanish at the end of inflation. Note added: After completing this work we have learnt [50] that our con- clusion that fNL is small for the specific double quadratic inflation model investigated in Sec. 5 is in qualitative agreement with an improved numeri- cal calculation by the authors of [22]. Acknowledgments: The authors are grateful to Kari Enqvist, Soo Kim, Andrew Liddle, David Lyth, Gerasimos Rigopoulos, Paul Shellard, Antti Vaihkonen and Bartjan van Tent for interesting discussions and David Lyth and Paul Shellard for comments on a draft of this paper. References [1] D. H. Lyth and A. Riotto, Phys. Rept. 314, 1 (1999) [arXiv:hep-ph/9807278]. [2] A. R. Liddle and D. H. Lyth, Cosmological inflation and large-scale structure (Cambridge University Press, Cambridge, England, 2000). [3] N. Bartolo, E. Komatsu, S. Matarrese and A. Riotto, Phys. Rept. 402, 103 (2004) [arXiv:astro-ph/0406398]. [4] D. Babich, P. Creminelli and M. Zaldarriaga, JCAP 0408, 009 (2004) [arXiv:astro-ph/0405356]. [5] J. Maldacena, JHEP 0305, 013 (2003) [arXiv:astro-ph/0210603]. [6] N. Arkani-Hamed, P. Creminelli, S. Mukohyama and M. Zaldarriaga, JCAP 0404, 001 (2004) [arXiv:hep-th/0312100]. [7] M. Alishahiha, E. Silverstein and D. Tong, Phys. Rev. D 70, 123505 (2004) [arXiv:hep-th/0404084]. [8] D. Seery and J. E. Lidsey, JCAP 0506, 003 (2005) [arXiv:astro-ph/0503692]. 26 [9] P. Creminelli, JCAP 0310, 003 (2003) [arXiv:astro-ph/0306122]. [10] D. H. Lyth, C. Ungarelli and D. Wands, Phys. Rev. D 67, 023503 (2003) [arXiv:astro-ph/0208055]. [11] F. Bernardeau and J. P. Uzan, Phys. Rev. D 66, 103506 (2002) [arXiv:hep-ph/0207295]; Phys. Rev. D 67, 121301 (2003) [arXiv:astro-ph/0209330]. [12] M. Zaldarriaga, Phys. Rev. D 69, 043508 (2004) [arXiv:astro-ph/0306006]. [13] K. Enqvist, A. Jokinen, A. Mazumdar, T. Multamaki and A. Vaihkonen, Phys. Rev. Lett. 94, 161301 (2005) [arXiv:astro-ph/0411394]. [14] N. Barnaby and J. M. Cline, arXiv:astro-ph/0601481. [15] A. Jokinen and A. Mazumdar, arXiv:astro-ph/0512368. [16] K. A. Malik and D. Wands, Class. Quant. Grav. 21, L65 (2004) [arXiv:astro-ph/0307055]. [17] G. I. Rigopoulos and E. P. S. Shellard, Phys. Rev. D 68 (2003) 123518 [arXiv:astro-ph/0306620]; [18] D. H. Lyth, K. A. Malik and M. Sasaki, JCAP 0505, 004 (2005) [arXiv:astro-ph/0411220]. [19] F. Vernizzi, Phys. Rev. D 71, 061301 (2005) [arXiv:astro-ph/0411463]. [20] D. H. Lyth and Y. Rodriguez, Phys. Rev. Lett. 95, 121302 (2005) [arXiv:astro-ph/0504045]. [21] D. Langlois and F. Vernizzi, Phys. Rev. Lett. 95, 091303 (2005) [arXiv:astro-ph/0503416]; Phys. Rev. D 72, 103501 (2005) [arXiv:astro-ph/0509078]. [22] G. I. Rigopoulos, E. P. S. Shellard and B. W. van Tent, arXiv:astro-ph/0511041; arXiv:astro-ph/0506704. [23] L. Alabidi and D. H. Lyth, arXiv:astro-ph/0510441. [24] K. Enqvist and A. Vaihkonen, JCAP 0409, 006 (2004) [arXiv:hep-ph/0405103]. [25] J. M. Bardeen, Phys. Rev. D 22, 1882 (1980). 27 [26] J. M. Bardeen, P. J. Steinhardt and M. S. Turner, Phys. Rev. D 28, 679 (1983). [27] D. H. Lyth and Y. Rodriguez, Phys. Rev. D 71, 123508 (2005) [arXiv:astro-ph/0502578]. [28] K. A. Malik, JCAP 0511, 005 (2005) [arXiv:astro-ph/0506532]. [29] D. Wands, K. A. Malik, D. H. Lyth and A. R. Liddle, Phys. Rev. D 62 (2000) 043527 [arXiv:astro-ph/0003278]. [30] A. A. Starobinsky, JETP Lett. 42, 152 (1985) [Pis. Hz. Esp. Tor. Fizz. 42, 124 (1985)]. [31] M. Sasaki and E. D. Stewart, Prog. Theor. Phys. 95, 71 (1996) [arXiv:astro-ph/9507001]. [32] M. Sasaki and T. Tanaka, Prog. Theor. Phys. 99, 763 (1998) [arXiv:gr-qc/9801017]. [33] D. H. Lyth and D. Wands, Phys. Rev. D 68, 103515 (2003) [arXiv:astro-ph/0306498]. [34] I. Zaballa, Y. Rodriguez and D. H. Lyth, arXiv:astro-ph/0603534. [35] D. Seery and J. E. Lidsey, JCAP 0509, 011 (2005) [arXiv:astro-ph/0506056]. [36] L. E. Allen, S. Gupta and D. Wands, JCAP 0601, 006 (2006) [arXiv:astro-ph/0509719]. [37] D. H. Lyth and I. Zaballa, arXiv:astro-ph/0507608. [38] J. Garcia-Bellido and D. Wands, Phys. Rev. D 52, 5636 (1995) [arXiv:gr-qc/9503049]. [39] D. Wands, N. Bartolo, S. Matarrese and A. Riotto, Phys. Rev. D 66, 043520 (2002) [arXiv:astro-ph/0205253]. [40] F. Bernardeau, L. Kofman and J. P. Uzan, Phys. Rev. D 70, 083004 (2004) [arXiv:astro-ph/0403315]. [41] D. H. Lyth, JCAP 0511, 006 (2005) [arXiv:astro-ph/0510443]. [42] M. P. Salem, Phys. Rev. D 72, 123516 (2005) [arXiv:astro-ph/0511146]. 28 [43] A. D. Linde, Phys. Rev. D 49, 748 (1994) [arXiv:astro-ph/9307002]. [44] E. J. Copeland, A. R. Liddle, D. H. Lyth, E. D. Stewart and D. Wands, Phys. Rev. D 49, 6410 (1994) [arXiv:astro-ph/9401011]. [45] D. Polarski and A. A. Starobinsky, Nucl. Phys. B 385, 623 (1992). [46] D. Langlois, Phys. Rev. D 59, 123512 (1999) [arXiv:astro-ph/9906080]. [47] C. Gordon, D. Wands, B. A. Bassett and R. Maartens, Phys. Rev. D 63, 023506 (2001) [arXiv:astro-ph/0009131]. [48] H. C. Lee, M. Sasaki, E. D. Stewart, T. Tanaka and S. Yokoyama, JCAP 0510, 004 (2005) [arXiv:astro-ph/0506262]. [49] W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numeri- cal Recipes in FORTRAN: The Art of Scientific Computing (Cambridge University Press, Cambridge, England, Second Edition, 1992). [50] E. P. S. Shellard, private communication (2006). [51] E. Komatsu and D. N. Spergel, Phys. Rev. D 63, 063002 (2001) [arXiv:astro-ph/0005036]. [52] E. Komatsu et al., Astrophys. J. Suppl. 148, 119 (2003) [arXiv:astro-ph/0302223]. [53] L. Verde, L. M. Wang, A. Heavens and M. Kamionkowski, Mon. Not. Roy. Astron. Soc. 313, L141 (2000) [arXiv:astro-ph/9906301]; 29
astro-ph/9408002
1
9408
1994-08-01T18:11:12
Dark Matter in the Light of COBE
[ "astro-ph" ]
The observations of all three COBE instruments are examined for the effects of dark matter. The anisotropy measured by the DMR, and especially the degree-scale ground- and balloon-based experiments, is only compatible with large-scale structure formation by gravity if the Universe is dominated by non-baryonic dark matter. The FIRAS instrument measures the total power radiated by cold dust, and thus places tight limits on the absorption of starlight by very cold gas and dust in the outer Milky Way. The DIRBE instrument measures the infrared background, and will place tight limits on the emission by low mass stars in the Galactic halo.
astro-ph
astro-ph
astro-ph/9408002, UCLA-ASTRO-ELW-94-01 DARK MATTER IN THE LIGHT OF COBE ∗ EDWARD L. WRIGHT UCLA Astronomy Los Angeles CA 90024-1562 ABSTRACT The observations of all three COBE instruments are examined for the effects of dark matter. The anisotropy measured by the DMR, and especially the degree- scale ground- and balloon-based experiments, is only compatible with large-scale structure formation by gravity if the Universe is dominated by non-baryonic dark matter. The FIRAS instrument measures the total power radiated by cold dust, and thus places tight limits on the absorption of starlight by very cold gas and dust in the outer Milky Way. The DIRBE instrument measures the infrared background, and will place tight limits on the emission by low mass stars in the Galactic halo. 1. Introduction While COBE (Boggess et al. 1992) has no instruments that directly detect dark matter, its three instruments offer important clues about the baryonic and non- baryonic content of the Universe. The FIRAS observations of the spectrum of the cosmic microwave background radiation (CMBR) show that any deviation from a blackbody are very small (Mather et al. 1990 and Mather et al. 1994). This limits the possible effect of energetic explosions on the formation of large-scale structure (Wright et al. 1994). If gravity is the force responsible for large-scale structure, then the DMR observations of anisotropy require a non-baryonic dark matter dominated Universe. Even the baryons in the Universe are mostly in a dark form, but FIRAS observations of the millimeter emission from the Galaxy show that these dark baryons can not be in clouds of very cold gas and dust associated with the CO absorbing clouds seen by Lequeux et al. (1993). However, even more compact configurations of baryons are allowed: brown dwarfs. A Galactic halo of old cold brown dwarfs will be essentially undetectable by the DIRBE instrument unless all of the mass is in objects right at the limit of hydrogen burning. ∗The National Aeronautics and Space Administration/Goddard Space Flight Center (NASA/GSFC) is responsible for the design, development, and operation of the Cosmic Background Explorer (COBE). Scientific guidance is provided by the COBE Science Working Group. GSFC is also responsible for the development of the analysis software and for the production of the mission data sets. Fig. 1.- Predicted ∆T for Holtzman models at 0.5◦ scale vs. quadrupole scale. 2. DMR ∆T and non-Baryonic Dark Matter The DMR anisotropy implies a small level of gravitational potential perturbations via the Sachs-Wolfe (1967) effect. At the same time, models of large-scale structure formation require certain levels of gravitational forces which can be converted into predicted ∆T 's. Figure 1 shows the predictions from the models of Holtzman (1989) RM Si0.5 and the anisotropy at 0.5◦ measured by compared to the COBE DMR hQ2 the MAX experiment (Clapp et al. 1994 and Devlin et al. 1994). The models with only baryonic matter are surrounded by open diamonds, while models emphasized by Wright et al. (1992) are surrounded by open circles. The CDM+baryon model and the vacuum dominated model (which still has 90% of the matter non-baryonic) both sit on top of the observed ∆T 's, while the open model and the mixed dark matter model need bias factors b8 < 2 to agree with the data. The nearest baryonic model needs b8 ≈ 10 to fit the data, which is not reasonable. This problem with baryonic models arises because non-baryonic dark matter perturbations start to grow at zeq ≈ 6000, while baryonic perturbations can only start to grow at zrec ≈ 103, and thus lose a factor of > ∼ 6 in growth. 8-14 /cm Flux in Microergs/cm^2/sr 470.00 486.00 502.00 518.00 534.00 550.00 Fig. 2.- The total FIRAS flux in the 8-14/cm band. 3. FIRAS Limits on Very Cold Dust Lequeux et al. (1993) have observed CO absorption toward extragalactic radio sources at low galactic latitudes. In at least one case the CO emission is very low, indicating very cold gas with Tex = 3.5 K. Since most surveys for interstellar material are based on measuring emission, there is the possibility that a large amount of very cold gas and dust could be hidden in cold clouds. But any starlight absorbed by these clouds will be reradiated in the millimeter region where the FIRAS instrument on COBE is sensitive. FIRAS uses bolometers which are approximately equally sensitive at all wavelengths, so the limit placed on the absorption of starlight by the cold clouds is roughly independent of their temperatures. Because of the measured Tex = 3.5 K, I have chosen the band 8 − 14 cm−1 which corresponds to hν/kTex = 3.3 to 5.8, which should include a fraction R 14 8 νβ [Bν(Td) − Bν(T◦)] dν R ∞ 0 νβ [Bν(Td) − Bν(T◦)] dν ≈ 0.5 (1) of the power power emitted by the very cold dust grains with emissivities varying like νβ for β ≈ 1.5 and dust temperature Td = 3.5 K. Lequeux et al. observe 4 clouds on a total path length P csc b = 66. Thus the optical depth from pole to pole of the disk is τ = 2 × 4/66 = 0.12. This implies that a 8-14 /cm Residual in Microergs/cm^2/sr -6.000 -2.000 2.000 6.000 10.000 14.000 Fig. 3.- The residual FIRAS flux after subtracting the monopole and dipole and fitting for Galactic dust. plane-parallel galactic disk in an isotropic optical background J will absorb a power per unit area of P = J Z sin b(1 − exp(−τ csc b)) cos bdbdl µ(1 − exp(−τ /µ))dµ = 1.24J. = 4πJ Z 1 0 (2) Considering only the part of the sky with sin b < 0.1, I expect that the flux per radian of Galactic plane observed by FIRAS will be approximately F = P ln(smax/smin) 4π (3) where the maximum distance along the line-of-sight smax can be taken as the exponential scale length of the disk, or about 3 kpc, while the minimum distance smin can be taken to be 10 times the scale height of the disk or about 1 kpc. Thus the very cold dust should radiate about 0.05J W/m2/rad into the 8-14 cm−1 band. The mean excess emission in the 8-14 cm−1 band over the model Iν(l, b) = Bν(T◦ + Dx cos l cos b + Dy sin l cos b + Dz sin b) + g(ν)G(l, b) (4) in the region with sin b < 0.1 and cos l < 0.866 is 6 × 10−10 W/m2/sr, where T◦ is the mean temperature of the cosmic background, Di are the components of the dipole anisotropy, g(ν) is the average galactic dust spectrum, and G(l, b) is the dust map (Wright et al. 1991). This model fixes the monopole and dipole using the high galactic latitude values, but allows for an adjustable amount of low latitude "normal" dust based on the 2-20 cm−1 spectrum (rather than using DIRBE or FIRAS data at higher frequencies), so there is only one parameter at each pixel. The fit removes some of the 8-14 cm−1 power but for dust with Td = 3.5 K and power law emissivities ∝ νβ with β between 0 and 1.8, the fraction of the total very cold dust emission within the 8-14 cm−1 band after the fit is 8 nνβ [Bν(Td) − Bν(T◦)] − Gg(ν)o dν R 14 R ∞ 0 νβ [Bν(Td) − Bν(T◦)] dν ≈ 0.3 with (5) (6) G = Pi νβ i [Bνi(Td) − Bνi(T◦)] g(νi)/σ2 i Pi g(νi)2/σ2 i using νi, g(νi) and σi from Mather et al. (1994). This reduces the expected excess to 0.03J W/m2/rad. Converting the observed excess to a flux per radian requires multiplying by the range of latitude used, which is ∆b = 0.2 rad. The result is 0.03J = (0.2 rad) × (6 × 10−10 W/m2/sr) = 1.2 × 10−11W/m2/rad (7) so I obtain J = 4 × 10−9 W/m2/sr. The total power in the 2.7 K background is 10−6 W/m2/sr so this limit is 0.4% of the CMB energy. The interstellar radiation field (ISRF) given by Wright (1993) integrates to 2.5 × 10−6 W/m2/sr, which is 600 times higher than J. Thus if I assume that the CO clouds seen by Lequeux et al. are optically thick to starlight, and that the local ISRF decays outward with the normal exponential disk scale length of 3 kpc, then the predicted millimeter excess from the Galactic plane in the 8-14 cm−1 range should be 600 times higher than it is. Even the heating provided by the diffuse extragalactic light (DEGL) would produce an 8-14 cm−1 excess that is 10 times higher than the observed excess. I conclude that CO clouds observed by Lequeux et al. can not be extremely optically thick objects hiding a large mass of very cold gas and dust. The observed number of clouds per unit csc b implies an absorption of the ISRF and DEGL that would produce a very large signal in the FIRAS Galactic plane observations, while only a small signal is seen. The low excitation temperature in CO is probably due to low density, giving T◦ ∼ Tex << Tgas ≈ Tdust. The clouds do absorb starlight, but the dust has a normal dust temperature, and the reradiation is already included in the FIRAS and DIRBE observations of the Galaxy. < DIRBE 3.5 Micron Flux 0.28 0.50 0.65 0.81 1.28 205.21 Fig. 4.- DIRBE flux in arbitrary linear units at 3.5 µm. The darkest part of the sky is ≈ 105 Jy/sr (Hauser 1994). 4. DIRBE Limits on a Brown Dwarf Halo Adams & Walker (1990) and Daly & McLaughlin (1992) have computed the expected intensity from a brown dwarf Galactic halo. The expected flux is generally very low unless all of the halo density is made up of maximum mass brown dwarfs, M = 0.08M⊙. The halo mass density assumed by Adams & Walker is ρ(r) = ρ◦ a2 a2 + r2 (8) with a = 2 kpc and ρ◦ = 0.19 M⊙/pc3 which gives a local density of 0.01 M⊙/pc3. The mass column density to the galactic pole is 137 M⊙/pc2 in this model. To convert this into a flux I need the mass and luminosity of a brown dwarf. I will assume M = 0.05 M⊙, and the Burrows, Hubbard & Lunine (1989) luminosity of 10−6 L⊙ with an effective temperature of 632 K at an age of 10 Gyr. With an M/L of 50,000 the resulting intensity at the pole is 137 M⊙/pc2/(4πM/L) = 2.2 × 10−4 L⊙/pc2/sr = 10−10 W/m2/sr. If I assume that the spectrum is a blackbody then the flux in individual bands is easy to find. The ratio νFν/Fbol = (15/π4)x4/(ex − 1) with x = hν/kT , which is 0.4 for x = 6.5 at 3.5 µm with T = 632 K. Thus the intensity is 48 Jy/sr in this model. Since the total brightness at the South Ecliptic Pole is 175 kJy/sr, the prospects for detecting an 0.03% effect due to brown dwarfs are not very good. Note, however, that the assumption of a blackbody spectrum is likely to be very bad. The spectrum of Jupiter shows very large features in the infrared. Warmer brown dwarfs with dusty atmospheres will have smoother spectra if the dust has a power law opacity κν ∝ νβ. Since dust absorbs more at short wavelengths than at long wavelengths, the color temperature of a dusty brown dwarf will be smaller than the effective temperature, leading to an apparent emissivity ǫ = (Tef f /Tcolor)4 that is close to 2. Since the emissivity is used to estimate the size of brown dwarfs, calculations based on blackbody emission will lead to sizes over-estimated by a factor of about 1.4. The lowered color temperature will further reduce the expected 3.5 µm intensity, and it also makes old cold brown dwarfs almost impossible to find in 2.2 µm surveys like those of Cowie et al. (1990). However, the individual brown dwarfs can easily be detected by the proposed Space InfraRed Telescope Facility (SIRTF). In a SIRTF 5′ × 5′ field of view, the closest brown dwarf will be 66 pc away, and produce a flux of 3.7 µJy at 3.5 µm, but only 0.32 µJy at 2.2 µm. The lowered color temperature expected for dusty brown dwarfs will give fluxes of 2.2 µJy at 3.5 µm, but only 0.09 µJy at 2.2 µm. SIRTF will have a sensitivity of 0.02 µJy per pixel at 3.5 µm in a 2500 second observation and will easily detect many brown dwarfs per FOV at 3.5 µm, even though the integrated intensity is too small to be seen by DIRBE. 5. Summary The observations by COBE of the CMBR show no evidence for non-gravitational forces producing large-scale structure. The gravitational forces implies by the DMR measurements of ∆T are sufficient to produce the observed large-scale structure only if most of the matter in the Universe responds freely to these gravitational forces before recombination, which requires non-baryonic dark matter. The baryonic dark matter cannot be very cold gas and dust associated with the CO absorption lines seen by Lequeux because it would produce too much millimeter wave emission from the Galactic plane. However, the dark baryons can easily be brown dwarfs which will escape detection by COBE and ground-based IR surveys but may well be seen by SIRTF. 6. References Adams, F. C. & Walker, T. P. 1990, ApJ, 359, 57-62. Boggess, N. W. et al. 1992, ApJ, 397, 420-429. Burrows, A., Hubbard, W. B & Lunine, J. I. 1989, ApJ, 345, 939-958. Clapp, A. C., Devlin, M. J., Gundersen, J. O., Hagmann, C. A., Hristov, V. V., Lange, A. E., Lim, M., Lubin, P. M., Mauskopf, P. D., Meinhold, P. R., Richards, P. L., Smoot, G. F., Tanaka, S. T., Timbie, P. T. & Wuensche, C. A. 1994, ApJL, TBD, TBD. Cowie, L. L., Gardner, J. P., Lilly, S. J. & McLean, I. 1990, ApJL, 360, L1-L4. Daly, R. A. & McLaughlin, G. C. 1992, ApJ, 390, 423-430. Devlin, M. J., Clapp, A. C., Gundersen, J. O., Hagmann, C. A., Hristov, V. V., Lange, A. E., Lim, M., Lubin, P. M., Mauskopf, P. D., Meinhold, P. R., Richards, P. L., Smoot, G. F., Tanaka, S. T., Timbie, P. T. & Wuensche, C. A. 1994, ApJL, 430, L1-L4. Hauser, M. G. 1994, presented at the Extragalactic Background Radiation Symposium at STScI, COBE Preprint 94-13. Holtzman, J. 1989, ApJSupp, 71, 1-24. Lequeux, J., Allen, R. J. & Guilloteau, S. 1993, A&A, 280, L23-L26. Mather, J. C., Cheng, E. S., Eplee, R. E., Jr., Isaacman, R. B., Meyer, S. S., Shafer, R. A., Weiss, R., Wright, E. L., Bennett, C. L, Boggess, N. W., Dwek, E., Gulkis, S., Hauser, M. G., Janssen, M., Kelsall, T., Lubin, P. M., Moseley, S. H. Jr., Murdock, T. L., Silverberg, R. F., Smoot, G. F. & Wilkinson, D. T. 1990, ApJ, 354, L37-L41. Mather, J. C., Cheng, E. S., Cottingham, D. A., Eplee, R. E., Jr., Fixsen, D. J., Hewagama, T., Isaacman, R. B., Jensen, K. A., Meyer, S. S., Noerdlinger, P. D., Read, S. M., Rosen, L. P., Shafer, R. A., Wright, E. L., Bennett, C. L., Boggess, N. W., Hauser, M. G., Kelsall, T., Moseley, S. H., Jr., Silverberg, R. F., Smoot, G. F., Weiss, R. & Wilkinson, D. T. 1994, ApJ, 420, 439-444. Wright, E. L. 1993, in "Back to the Galaxy", AIP Conference Proceedings 278, eds. S. S. Holt & F. Trimble, p. 193. Wright, E. L. etal. 1992, ApJL, 396, L13. Wright, E. L., Mather, J. C., Bennett, C. L., Cheng, E. S., Shafer, R. A., Fixsen, D. J., Eplee, R. E. Jr., Isaacman, R. B., Read, S. M., Boggess, N. W., Gulkis, S. G., Hauser, M. G., Janssen, M., Kelsall, T., Lubin, P. M., Meyer, S. S., Moseley, S. H. Jr., Murdock, T. L., Silverberg, R. F., Smoot, G. F., Weiss, R., and Wilkinson, D. T., 1991, ApJ, 381, 200-209. Wright, E. L., Mather, J. C., Fixsen, D. J., Kogut, A., Shafer, R. A., Bennett, C. L., Boggess, N. W., Cheng, E. S., Silverberg, R. F., Smoot, G. F. & Weiss, R. 1994a, ApJ, 420, 450.
astro-ph/0011250
1
0011
2000-11-13T17:00:47
Measuring Stellar and Dark Mass Fractions in Spiral Galaxies
[ "astro-ph" ]
We explore the relative importance of the stellar mass density as compared to the inner dark halo, for the observed gas kinematics thoughout the disks of spiral galaxies. We perform hydrodynamical simulations of the gas flow in a sequence of potentials with varying the stellar contribution to the total potential. The stellar portion of the potential was derived empirically from K-band photometry. The output of the simulations - namely the gas density and the gas velocity field - are then compared to the observed spiral arm morphology and the H-alpha gas kinematics. We solve for the best matching spiral pattern speed and draw conclusions on how massive the stellar disk can be at most. For the case of the galaxy NGC 4254 (Messier 99) we demonstrate that the prominent spiral arms of the stellar component would overpredict the non-circular gas motions unless an axisymmetric dark halo component adds significantly in the radial range R_exp < R < 3*R_exp.
astro-ph
astro-ph
Measuring Stellar and Dark Mass Fractions in Spiral Galaxies Thilo Kranz, Adrianne Slyz, and Hans-Walter Rix Max-Planck-Institut fur Astronomie, Konigstuhl 17, 69117 Heidelberg, Germany Abstract. We explore the relative importance of the stellar mass density as compared to the inner dark halo, for the observed gas kinematics thoughout the disks of spiral galaxies. We perform hydrodynamical simulations of the gas flow in a sequence of po- tentials with varying the stellar contribution to the total potential. The stellar portion of the potential was derived empirically from K-band photometry. The output of the simulations – namely the gas density and the gas velocity field – are then compared to the observed spiral arm morphology and the Hα gas kinematics. We solve for the best matching spiral pattern speed and draw conclusions on how massive the stellar disk can be at most. For the case of the galaxy NGC 4254 (Messier 99) we demonstrate that the prominent spiral arms of the stellar component would overpredict the non-circular gas motions unless an axisymmetric dark halo component adds significantly in the radial range Rexp < R < 3 Rexp. 1 Introduction In almost all galaxy formation scenarios, non-baryonic dark matter plays an im- portant role. Today's numerical simulations of cosmological structure evolution quite successfully reproduce the observed galaxy distribution in the universe [3]. While galaxies form and evolve inside dark halos their physical appearance depends strongly on the local star formation and merging history. At the same time the halos evolve and merge as well. According to the simulations, we expect that the dark matter is important in the inner parts of galaxies [5],[6] and that it thus has a considerable influence on the kinematics. These predictions are in contrast to some studies which indicate that galactic stellar disks - at least of barred spiral galaxies - alone dominate the kinematics of the inner regions [1]. Apparently this is also the case in our own Milky Way [2]. Determining individual mass fractions of the luminous and dark matter is not a straightforward task. The rotation curve of a disk galaxy is only sensitive to the total amount of gravitating matter, but does not allow the distinction between the two mass density profiles. For a detailed analysis it is necessary to adopt more refined methods to separate out the different profiles. Previous investigations used for example knowledge of the kinematics of rotating bars [8] or the geometry of gravitational lens systems [4]. Here we would like to ex- ploit the fact, that the stellar mass in disk galaxies is often organized in spiral arms, i.e. in coherent non-axisymmetric structures. In most proposed scenarios, the dark matter, however, is collisionless and dominated by random motions. Therefore it is not susceptible to spiral structures. If the stellar mass dominates, 2 Kranz et al. Fig. 1. Near infrared K'-band (2.1 m m) image of NGC 4254 with a total exposure time of 20 minutes at the Calar Alto 3.5 m telescope. Bright foreground stars are masked out. The overlay shows the slit orientations of the spectrograph. We took 8 longslit spectra (angles labeled in bold font) crossing the galaxy's center to measure the 2D velocity field the spiral arms, as traced by the near infrared (NIR) light, should induce con- siderable non-circular motions in the gas, that manifest themselves as velocity "wiggles" in observed gas kinematics. Using hydrodynamical gas simulations we are able to predict these velocity wiggles and compare them to the observations. Hence the contribution of the perturbative forces with respect to the total forces can be determined quantitatively and can be used to constrain the stellar disk to halo mass ratio. 2 Observations For this analysis we need data to provide us with information on the stellar mass distribution and on the gas kinematics of a sample of galaxies. To map the stellar surface mass density it is most desirable to take NIR images of the galaxies, because in this waveband dust extinction and population effects are minimized [7]. During two observing runs in May 1999 and March 2000, we obtained photometric data for 20 close-by NGC galaxies. We used the Omega Prime camera at the Calar Alto 3.5 m telescope with the K-band filter (K' at Stellar and Dark Mass Fractions in Spirals 3 Fig. 2. Simulation results for the gas density. Here we depict the best fitting result (in contours) overlaid on a deprojected image of NGC 4254. To enhance the contrast of the spiral arms, an axisymmetric model of the galaxy has been subtracted from the original image. We find an excellent match to the spiral morphology 2.1 m m). It provides us with a field of view of 6.′76 x 6.′76. Figure 1 shows the K-band image of the Messier galaxy M99. The labeled one arcminute scale bar translates to 5.8 kpc within the galaxy. The kinematic data were obtained with the TWIN, a longslit spectrograph at the 3.5 m telescope. For a reasonable coverage of a galaxy's velocity field, we needed to take 8 slit positions (or 16 position angles) across the entire disk of the galaxy (also displayed in Figure 1). We chose longslit-spectroscopy rather than fabry-perot interferometry because of its higher spectral resolution and better sensitivity to faint emission. So far we were able to collect complete sets of longslit spectra for only four galaxies, mostly due to only moderate weather conditions during the spectroscopy runs. 3 First results As a pilot project, we analyzed the data of NGC 4254 (M99), a late type spi- ral galaxy with strong arms, located in the Virgo cluster. Assuming a constant stellar mass-to-light ratio in the K-band image, the gravitational potential due to the stellar mass fraction was calculated by direct integration over the whole 4 Kranz et al. Fig. 3. Simulation results of the gas velocity. Here we compare two simulations (con- tinuous lines) with different disk mass fractions fd = Mdisk/Mmaxdisk to the observed kinematics (data points). The maximal disk case (right) is clearly not matching the observations well. Displayed is only one (135◦) out of 16 slit position angles light/mass distribution. The mass-to-light ratio for the maximum disk contribu- tion was scaled by the measured rotation curve. For the dark matter contribution we assumed an isothermal halo with a core. To combine the two components we chose a stellar mass fraction fd = Mdisk/Mmaxdisk and added the halo with the variable parameters adjusted to give a best fit to the rotation curve. We used this potential as an input for the hydrodynamical gas simulations. Figure 2 presents the resulting gas surface density, as it settles in the poten- tial. The morphology of the gas distribution is very sensitive to the speed with which the spiral pattern of the galaxy rotates (pattern speed). In figure 2 we printed the result of the simulation whose spiral structure best matches the K- band image morphology. We find quite good agreement for a pattern speed of Ωp = 23 km s−1 kpc−1, which places the corotation radius to 8.4 kpc. To quantify the relevance of the stellar mass component, we need now to fit the velocities – in particular the amplitude of the wiggles. In figure 3 a comparison of the modelled and the measured rotation curves are presented for the position angle of 135◦. The two panels show the rotation curves for different disk mass fractions: less than half maximal and the maximum disk case. 4 Discussion and outlook Although there is quite some scatter in the observed gas kinematics, we find that the velocity jumps, which are apparent in the simulations for the maximum disk case are too large to be in agreement with the measurements. The inner part of the simulated rotation curve (< 0.′3) is dominated by the dynamics of the small bar, which is present at the center of the galaxy. Its pattern speed might be different from the one of the spiral's and thus relate to a mismatch in the Stellar and Dark Mass Fractions in Spirals 5 inner part of the rotation curve. We conclude that an axisymmetric dark halo is needed to explain the kinematics of the stellar disk. The influence of the stellar disk is submaximal in the sense that we don't find strong enough velocity wiggles in the observed kinematics as would be expected if the stellar disk was the major gravitating source inside the inner few disk scale lengths. How this conclusion might apply to other spiral galaxies will be the upcoming issue of this project. We plan to extend our analysis at first to the 3 other galaxies where we have already now complete data sets. Finally we intend to draw our final conclusions on a basis of a sample consisting of 8 - 10 members. This should be sufficient to determine reliable results about the luminous and dark mass distributions in spiral galaxies. References 1. V. P. Debattista, J. A. Sellwood: ApJ in press, astro-ph/0006275 (2000) 2. O. Gerhard: 'Dynamics of the Galaxy'. In: Galaxy Dynamics, Rutgers University, USA, August 8–12, 1998, ed. by D. Merritt, J. A. Sellwood, M. Valluri (ASP Con- ference Series, Vol 182, 1999), pp. 307–320 3. G. Kauffmann, J. Colberg, A. Diaferio, S. D. M. White: MNRAS, 303, 188 (1999) 4. A. Maller et al.: ApJ, 533, 194 (2000) 5. J. Navarro, C. Frenk, S. D. White: ApJ, 462, 563 (1996) 6. J. Navarro, C. Frenk, S. D. White: ApJ, 490, 493 (1997) 7. H.-W. Rix, M. Rieke: ApJ, 418, 123 (1993) 8. B. J. Weiner, J. A. Sellwood, T. B. Williams: ApJ in press, astro-ph/0008205 (2000)
astro-ph/0507207
1
0507
2005-07-08T13:01:22
Observations of Selected AGN with H.E.S.S
[ "astro-ph" ]
A sample of selected active galactic nuclei (AGN) was observed in 2003 and 2004 with the High Energy Stereoscopic System (H.E.S.S.), an array of imaging atmospheric-Cherenkov telescopes in Namibia. The redshifts of these candidate very-high-energy (VHE, >100 GeV) gamma-ray emitters range from z=0.00183 to z=0.333. Significant detections were already reported for some of these objects, such as PKS 2155-304 and Markarian 421. Marginal evidence (3.1 sigma) for a signal is found from large-zenith-angle observations of Markarian 501, corresponding to an integral flux of I(>1.65 TeV) = (1.5 +/- 0.6 (stat) +/- 0.3 (syst)) x 10^{-12} cm^{-2} s^{-1} or ~15% of the Crab Nebula flux. Integral flux upper limits for 19 other AGN, based on exposures of ~1 to ~8 hrs live time, and with average energy thresholds between 160 GeV and 610 GeV, range from 0.4% to 5.1% of the Crab Nebula flux. All the upper limits are the most constraining ever reported for these objects.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. agnul (DOI: will be inserted by hand later) December 1, 2018 Observations of Selected AGN with H.E.S.S. F. Aharonian1, A.G. Akhperjanian2, A.R. Bazer-Bachi3, M. Beilicke4, W. Benbow1, D. Berge1, K. Bernlohr1 , 5, C. Boisson6, O. Bolz1, V. Borrel3, I. Braun1, F. Breitling5, A.M. Brown7, P.M. Chadwick7, L.-M. Chounet8, R. Cornils4, L. Costamante1 , 20, B. Degrange8, H.J. Dickinson7, A. Djannati-Atai9, L.O'C. Drury10, G. Dubus8, D. Emmanoulopoulos11, P. Espigat9, F. Feinstein12, G. Fontaine8, Y. Fuchs13, S. Funk1, Y.A. Gallant12, B. Giebels8, S. Gillessen1, J.F. Glicenstein14, P. Goret14, C. Hadjichristidis7, M. Hauser11, G. Heinzelmann4, G. Henri13, G. Hermann1, J.A. Hinton1, W. Hofmann1, M. Holleran15, D. Horns1, A. Jacholkowska12, O.C. de Jager15, B. Kh´elifi1, Nu. Komin5, A. Konopelko1 5, I.J. Latham7, , R. Le Gallou7, A. Lemi`ere9, M. Lemoine-Goumard8, N. Leroy8, T. Lohse5, J.M. Martin6, O. Martineau-Huynh16, A. Marcowith3, C. Masterson1 , 20, T.J.L. McComb7, M. de Naurois16, S.J. Nolan7, A. Noutsos7, K.J. Orford7, J.L. Osborne7, M. Ouchrif16 , 20, M. Panter1, G. Pelletier13, S. Pita9, , 11, M. Punch9, B.C. Raubenheimer15, M. Raue4, J. Raux16, S.M. Rayner7, A. Reimer17, G. Puhlhofer1 O. Reimer17, J. Ripken4, L. Rob18, L. Rolland16, G. Rowell1, V. Sahakian2, L. Saug´e13, S. Schlenker5, R. Schlickeiser17, C. Schuster17, U. Schwanke5, M. Siewert17, H. Sol6, D. Spangler7, R. Steenkamp19, C. Stegmann5, J.-P. Tavernet16, R. Terrier9, C.G. Th´eoret9, M. Tluczykont8 C. Venter15, P. Vincent16, H.J. Volk1, and S.J. Wagner11 , 20, G. Vasileiadis12, 1 Max-Planck-Institut fur Kernphysik, Heidelberg, Germany; 2 Yerevan Physics Institute, Armenia; 3 Centre 4 Universitat Hamburg, Institut fur d'Etude Spatiale des Rayonnements, CNRS/UPS, Toulouse, France; Experimentalphysik, Germany; 5 Institut fur Physik, Humboldt-Universitat zu Berlin, Germany; 6 LUTH, UMR 8102 du CNRS, Observatoire de Paris, Section de Meudon, France; 7 University of Durham, Department 8 Laboratoire Leprince-Ringuet, IN2P3/CNRS, Ecole Polytechnique, Palaiseau, France; of Physics, U.K.; 9 APC, Paris, France; ⋆ 10 Dublin Institute for Advanced Studies, Ireland; 11 Landessternwarte, Konigstuhl, Heidelberg, Germany; 12 Laboratoire de Physique Th´eorique et Astroparticules, IN2P3/CNRS, Universit´e Montpellier II, France; 13 Laboratoire d'Astrophysique de Grenoble, INSU/CNRS, Universit´e Joseph Fourier, France; 14 DAPNIA/DSM/CEA, CE Saclay, Gif-sur-Yvette, France; 15 Unit for Space Physics, North- West University, Potchefstroom, South Africa; 16 Laboratoire de Physique Nucl´eaire et de Hautes Energies, IN2P3/CNRS, Universit´es Paris VI & VII, France; 17 Institut fur Theoretische Physik, Lehrstuhl IV, Ruhr- Universitat Bochum, Germany; 18 Institute of Particle and Nuclear Physics, Charles University, Prague, Czech Republic; 19 University of Namibia, Windhoek, Namibia; 20 European Associated Laboratory for Gamma-Ray Astronomy, jointly supported by CNRS and MPG; Received 20 May 2005; Accepted 16 June 2005 Abstract. A sample of selected active galactic nuclei (AGN) was observed in 2003 and 2004 with the High Energy Stereoscopic System (H.E.S.S.), an array of imaging atmospheric-Cherenkov telescopes in Namibia. The redshifts of these candidate very-high-energy (VHE, >100 GeV) γ-ray emitters range from z=0.00183 to z=0.333. Significant detections were already reported for some of these objects, such as PKS 2155−304 and Markarian 421. Marginal evidence (3.1σ) for a signal is found from large-zenith-angle observations of Markarian 501, corresponding to an integral flux of I(>1.65 TeV) = (1.5±0.6stat ±0.3syst) × 10−12 cm−2 s−1 or ∼15% of the Crab Nebula flux. Integral flux upper limits for 19 other AGN, based on exposures of ∼1 to ∼8 hrs live time, and with average energy thresholds between 160 GeV and 610 GeV, range from 0.4% to 5.1% of the Crab Nebula flux. All the upper limits are the most constraining ever reported for these objects. Key words. Galaxies: active - BL Lacertae objects: Individual - Gamma rays: observations ⋆ UMR 7164 (CNRS, Universit´e Paris VII, CEA, Observatoire de Paris) Active galactic nuclei are known to emit radiation over the entire electromagnetic spectrum, from radio waves to TeV γ-rays. These objects, which are found in only a small 1. Introduction 2 F. Aharonian et al.: Observations of Selected AGN with H.E.S.S. fraction of the total number of observed galaxies, are very luminous, extremely compact, and can exhibit large lu- minosity variations on time scales ranging from less than an hour up to several years. Although AGN differ widely in their observed characteristics, a unified description (as reviewed in Urry & Padovani 1995) has emerged in which an AGN consists of a super-massive black hole (107 − 1010 solar masses) surrounded in the inner regions by an ac- cretion disk, and in the outer regions by a thick torus of gas and dust. In some AGN (the radio-loud popula- tion, ∼10%), a highly relativistic outflow of energetic par- ticles exists approximately perpendicular to the accretion disk and torus plane. This flow forms collimated radio- emitting jets which generate the non-thermal emission ob- served from radio to γ-rays. It is believed that some of the numerous AGN classifications result from viewing these objects at various orientation angles with respect to the torus plane. Essentially all AGN detected at VHE ener- gies (shown in order of redshift with references in Table 1) are radio-loud objects of the BL Lacertae (BL Lac) type, which have their jet pointed close to the observer's line of sight. An exception to this exists with the detection of the Fanaroff-Riley type I radio galaxy M 87 at VHE ener- gies. Although M 87 is believed to be a mis-aligned BL Lac (Tsvetanov et al. 1998), it is not clear whether the VHE emission comes from the jet or the central object. The known VHE AGN have helped to constrain signif- icantly the models for production of VHE γ-rays through spectral and variability studies. However, there are still many differing models that describe the present data, making a larger sample of known VHE AGN necessary to make more definitive conclusions. Also, VHE photons are absorbed by interactions on the extragalactic back- ground light (EBL) leading to an energy dependent hori- zon for viewing VHE sources. The energy spectrum of VHE AGN may exhibit characteristics, such as steepen- ing of the spectrum and a cutoff, as a result of this ab- sorption. Interpretation of such features can be used as a probe of the EBL in the optical and near-IR regimes (Stecker, de Jager & Salamon 1992; Schroedter 2005), for which direct measurements are dominated by large sys- tematic uncertainties. Since such an interpretation is com- plicated by discerning which features are a result of EBL absorption and which are intrinsic to the object, a large data set of VHE AGN at differing redshifts are needed to ascertain which effects can be attributed to the EBL. A large sample of AGN located at z<0.333 was observed by H.E.S.S. in 2003 and 2004. Most of these objects are BL Lacs, many of which are sug- gested as good candidates for detection as VHE emitters (Costamante & Ghisellini 2002; Perlman 1999; Stecker, de Jager & Salamon 1996). A sample of nearby non-blazar AGN, like M 87, was also observed with the hope of extending the known VHE-bright AGN to other classes. These include a set of famous radio-loud galax- ies, characterized by resolved radio, optical and X-ray jets (Cen A, Pictor A, 3C 120, and the quasar 3C 273) and a sample of radio-weak objects (the Seyfert galaxies NGC 1068, NGC 3783 and NGC 7469). The detections re- sulting from the H.E.S.S. AGN observation program have been reported elsewhere (see Table 1 for references). These include the confirmation of the VHE emission seen from M 87 and PKS 2155−304, the detection of Markarian 421 using large-zenith-angle observations, and the discovery of VHE emission from PKS 2005−489. Flux upper limits, the strongest ever produced, from the non-detection of the remaining objects are presented here. 2. H.E.S.S. Detector The H.E.S.S. experiment, a square array (120 m side) of four imaging atmospheric-Cherenkov telescopes located in the Khomas Highland of Namibia (23◦ 16' 18" S, 16◦ 30' 1" E, 1800 m above sea level), uses stereoscopic obser- vations of γ-ray induced air showers to search for astro- physical γ-ray emission above ∼100 GeV. Each telescope has a 107 m2 tessellated mirror dish and a 5◦ field-of-view (f.o.v.) camera consisting of 960 individual photomulti- plier pixels. The sensitivity of H.E.S.S. (5σ in 25 hours for a 1% Crab Nebula flux source at 20◦ zenith angle) allows for detection of VHE emission from objects at previously undetectable flux levels. More details on H.E.S.S. can be found in Bernlohr et al. (2003), Funk et al. (2004), Hofmann (2003), and Vincent et al. (2003). 3. Observations The H.E.S.S. observations of AGN in 2004 use the full four-telescope array. For some of the data, individual tele- scopes were excluded from the observations or analysis due to hardware problems. Also, 2003 observations of 1ES 0323+022 were made prior to the completion of the array and thus use only two or three telescopes. While the sensitivity of H.E.S.S. is less during observations with fewer telescopes, it is still unprecedented. Table 2 shows the candidate AGN observed by H.E.S.S. and gives details of the observations that pass selection criteria which re- move data for which the weather conditions were poor or the hardware was not functioning properly. The data were taken in 28 minute runs using Wobble mode, i.e. the source direction is offset, typically by ±0.5◦, relative to the cen- ter of the f.o.v. of the camera during observations, which allows for both on-source observations and simultaneous estimation of the background induced by charged cosmic rays. As the energy threshold of H.E.S.S. observations in- creases with zenith angle, the mean zenith angle of the ex- posure for each of the AGN along with the corresponding average energy threshold (after selection cuts) of those ob- servations is also shown in Table 2. It should be noted that the H.E.S.S. Monte Carlo simulations show that the az- imuthal angle at which an object is observed has a small ef- fect on the energy threshold of observations. Sources which culminate in the south (i.e. those with declination less than the latitude of H.E.S.S.) have slightly higher energy thresholds (e.g. compare 1ES 0323+022 with Pictor A). F. Aharonian et al.: Observations of Selected AGN with H.E.S.S. 3 Table 1. Known VHE AGN in order of redshift, along with the references for the initial discovery by other VHE instruments and references for H.E.S.S. results. Known is defined here as having VHE detections reported by at least two different instruments or by H.E.S.S.. AGN M 87 Markarian 421 Markarian 501 1ES 2344+514 1ES 1959+650 PKS 2005-489 PKS 2155-304 1ES 1426+428 z 0.004 0.030 0.034 0.044 0.047 0.071 0.116 0.129 Reference H.E.S.S. Reference Aharonian et al. (2003) Beilicke et al. (2005) Punch et al (1992) Quinn et al (1996) Catanese et al. (1998) Nishiyama et al.(1999) Chadwick et al. (1999a) Horan et al. (2002) Aharonian et al. (2005c) Aharonian et al. (2005b) Aharonian et al. (2005a) 4. Analysis Technique The data passing the run selection criteria are calibrated as detailed in Aharonian et al. (2004b), and the event re- construction and background rejection are performed as described in Aharonian et al. (2005a), with some minor improvements discussed in Aharonian et al. (2005b). The background is estimated using all events passing selec- tion cuts in a number of circular off-source regions offset by the same distance, relative to the center of the f.o.v., in the sky as the on-source region (for more details see Aharonian et al. 2001). The on-source region, the size of which is optimized for the detection of point sources, is a circle of radius ∼0.11◦ centered on the source, and each off- source region has approximately1 the same area as the on- source region. The maximum number of non-overlapping off-source regions fitting in the field of view are used. An area around the on-source position, completely containing the H.E.S.S. point-spread-function, is excluded to elim- inate possible contamination from poorly reconstructed γ-rays. For the typical on-source offset of 0.5◦, eleven off- source regions are possible. In the case of observations of Cen A, offset by 0.7◦, sixteen regions are used. The sta- tistical error on the background measurement is reduced by the use of a larger background sample, and there is no need for a radial acceptance correction, which accounts for the strongest acceptance change across the f.o.v., since the off-source regions are offset by the same radial dis- tance as the on-source region. The significance of any ex- cess is calculated following the method of Equation (17) in Li & Ma (1983) and all upper limits are determined using the method of Feldman & Cousins (1998). 5. Results Figure 1 shows the distribution of the significance ob- served from the direction of each of the twenty AGN. No significant excess of VHE γ-rays is found from any of the AGN in the given exposure time (<8 hrs each), with the possible exception of Markarian 501 (3.1σ). Specific de- 1 The off-source data are first placed into a pixelated two- dimensional map and then integrated in an approximate circle for each region. The difference in total area is of order 1%. s e i r t n E 5 4 3 2 1 0 -4 -3 -2 -1 0 1 2 3 4 σ Fig. 1. Distribution of the significance observed from the 20 selected AGN. The curve represents a Gaussian distri- bution with zero mean and a standard deviation of one. tails of the results for each AGN are shown in Table 2. Additionally, a search for serendipitous source discoveries in the H.E.S.S. f.o.v. centered on each of the AGN yielded no significant excess. Given that it is well established that Markarian 501 is a VHE γ-ray emitter, the excess (3.1σ) from the only night (MJD 53172) of observations of Markarian 501 can be treated as significant and a flux calculated. Assuming the spectrum measured above 1.5 TeV by the High Energy Gamma Ray Astronomy (HEGRA) experiment (Bradbury et al. 1997), a power law with a photon index of Γ=2.6, the corresponding integral flux above the 1.65 TeV energy threshold is I(>1.65 TeV) = (1.5±0.6stat±0.3syst) × 10−12 cm−2 s−1 or ∼15% of the H.E.S.S. Crab Nebula flux above this threshold. While the VHE flux from Markarian 501 is known to be highly vari- able, the measured flux is similar to the value reported in Bradbury et al. (1997). For the remaining undetected AGN, 99.9% upper lim- its on the integral flux (assuming a power law spec- trum with Γ=3.0) above the energy threshold of the ob- servations, and references to previously published limits (when available), are shown in Table 3. The photon index, Γ=3.0, was chosen for two reasons: First, the recently mea- Table 2. The candidate AGN ordered by right ascension in groups of BL Lacs, other radio-loud galaxies and radio-weak galaxies. The coordinates (J2000), redshift, and type (BL=BL Lacs, FSRQ=Flat Spectrum Radio Quasar, Sy=Seyferts (types I & II), FR=Fanaroff-Rileys (types I & II)) shown are taken from the SIMBAD Astronomical Database and the NASA/IPAC Extragalactic Database. Observations not using the full array are shown by the number of telescopes (Ntel) column along the breakdown of the number of observation runs (Nruns) and live time (T). Ntel=2∗ refers to a configuration of two telescopes on opposite sides of the square (170 m separation). The mean zenith angle of observation (Zobs), the corresponding post-selection cuts energy threshold (Eth), the number of on-source and off-source counts, the off-source normalization, the observed excess and significance (S) are also shown. Object α δ z Type Ntel Nruns [hh mm ss] [dd mm ss] T [hrs] Zobs [◦] Eth [GeV] On Off Norm Excess S [σ] BL Lacs 1ES 0145+138 1ES 0229+200 1ES 0323+022 PKS 0548−322 EXO 0556.4−3838 RGB J0812+026 RGB J1117+202 1ES 1440+122 Markarian 501 RBS 1888 Q J22548−2725 PKS 2316−423 1ES 2343−151 Radio-Loud Galaxies 01 48 29.8 02 32 48.6 03 26 14.0 05 50 40.6 05 58 06.2 08 12 01.9 11 17 06.2 14 42 48.3 16 53 52.2 22 43 42.0 22 54 53.2 23 19 05.9 23 45 37.8 +14 02 19 +20 17 17 +02 25 15 −32 16 16 −38 38 27 +02 37 33 +20 14 07 +12 00 40 +39 45 37 −12 31 06 −27 25 09 −42 06 49 −14 49 10 0.125 0.140 0.147 0.069 0.034 -- 0.139 0.162 0.034 0.226 0.333 0.055 0.226 BL BL BL BL BL BL BL BL BL BL BL BL BL 3C 120 Pictor A 3C 273 Cen A 04 33 11.1 05 19 49.7 12 29 06.7 13 25 27.6 +05 21 16 −45 46 45 +02 03 09 −43 01 09 0.033 0.034 0.158 0.00183 FR I FR II FSRQ FR I Radio-Weak Galaxies NGC 1068 NGC 3783 NGC 7469 02 42 40.8 11 39 01.8 23 03 15.8 −00 00 48 −37 44 19 +08 52 26 0.00379 0.00965 0.016 Sy II Sy I Sy I 3,4 4 12 (4,8) 4.3 (1.5,2.8) 2 0.8 2∗,3,4 12 (3,3,6) 4.7 (1.3,1.1,2.3) 4 4 3,4 3,4 4 4 4 4 4 3,4 4 4 4 4 4 11 3 4 (2,2) 9 (3,6) 4.1 1.2 0.7 (0.5,0.2) 3.8 (1.2,2.6) 2 4 6 5 5 0.9 1.8 2.6 2.1 2.2 6 (2,4) 2.6 (0.9,1.7) 12 18 9 10 10 5.0 7.4 3.9 4.2 4.3 2∗,3,4 5 (1,2,2) 1.8 (0.4,0.8,0.8) 4 10 4.3 40 46 29 20 33 30 52 38 64 15 13 21 11 32 27 37 21 27 27 33 310 410 210 190 250 220 610 290 1650 170 170 190 160 230 220 280 190 210 220 250 206 24 279 340 101 44 155 59 95 190 200 169 235 232 516 245 465 246 106 271 2287 280 3009 3321 934 519 1703 633 742 2330 2294 2032 2652 3023 5451 2634 7065 1.2 −1.4 7.2 39.0 17.2 −2.7 1.9 2.3 28.1 0.1 0.08953 −0.3 0.09077 0.4 0.09034 2.1 0.09065 1.7 0.08968 −0.4 0.08999 0.1 0.08990 0.3 0.08955 0.09017 3.1 0.08958 −18.7 −1.3 0.08986 −0.4 0.08952 −12.9 −0.9 −0.2 0.08966 −6.1 −2.8 0.09087 −42.7 −2.5 1.0 0.09041 0.5 0.08978 0.06461 0.4 23.2 8.5 8.5 2580 1205 3373 0.8 0.09047 0.08947 −0.2 0.08962 −31.2 −1.8 12.6 −1.8 4 F . A h a r o n i a n e t a l . : O b s e r v a t i o n s o f S e l e c t e d A G N w i t h H E . S . S . . F. Aharonian et al.: Observations of Selected AGN with H.E.S.S. 5 Table 3. Integral flux upper limits (99.9% confidence level, assuming a power law spectrum with Γ=3.0) above the energy threshold of observations (Eth) and the corresponding percentage of the H.E.S.S. Crab Nebula flux from observations of each of the candidate AGN by H.E.S.S.. References to upper limits available from other VHE instru- ments are also shown: HEGRA (Aharonian et al. 2004a), Whipple (a: Buckley 1999; b: de la Calle Perez et al. 2003; c: Horan et al. 2004), CANGAROO (a: Roberts et al. 1998; b: Roberts et al. 1999; c: Rowell et al. 1999, d: Nishijima 2002), University of Durham Mark VI Telescope (a: Chadwick et al. 1999b; b: Chadwick et al. 2000), and Milagro (Williams 2004). Object z Eth [GeV] I(>Eth) [10−12 cm−2 s−1] Crab % Previous Observations BL Lacs 1ES 0145+138 1ES 0229+200 1ES 0323+022 PKS 0548−322 EXO 0556.4−3838 RGB J0812+026 RGB J1117+202 1ES 1440+122 RBS 1888 Q J22548−2725 PKS 2316−423 1ES 2343−151 Radio-Loud Galaxies 3C 120 Pictor A 3C 273 Cen A Radio-Weak Galaxies NGC 1068 NGC 3783 NGC 7469 0.125 0.140 0.147 0.069 0.034 -- 0.139 0.162 0.226 0.333 0.055 0.226 0.033 0.034 0.158 0.00183 0.00379 0.00965 0.016 310 410 210 190 250 220 610 290 170 170 190 160 230 220 280 190 210 220 250 2.11 2.76 3.92 6.65 10.1 7.49 1.44 5.11 3.19 5.83 4.13 6.43 0.92 3.33 2.90 5.68 3.28 6.04 1.27 HEGRA, Whipplec HEGRA, Whippleb,c, Milagro 1.5 3.1 HEGRA, Whipplec, Mark VIa, Milagro 1.5 2.2 Whipplea, CANGAROOa,b,d, Mark VIb 5.1 3.1 3.0 3.3 0.9 1.6 1.4 1.6 Whipplea, CANGAROOa, Mark VIa HEGRA, Whippleb, Milagro Milagro HEGRA, Milagro HEGRA HEGRA, Whipplea CANGAROOc, Mark VIa 0.4 1.4 1.8 1.9 1.3 2.5 0.6 sured VHE spectra of several AGN (e.g. PKS 2155−304, PKS 2005−489) are softer than the Crab Nebula-like index of Γ=2.5 often used for VHE upper limits in past publica- tions. Second, the softer index was chosen to account for the possible steepening of the observed spectra of the AGN due to the absorption of γ-rays on the EBL. Assuming a different photon index (i.e. Γ between 2.5 and 3.5) has less than a ∼10% effect on the reported limits, and the system- atic error on the upper limits is estimated to be ∼20%. The percentage of the Crab Nebula flux shown in Table 3 is cal- culated relative to the integral flux, above the same thresh- old, determined from the H.E.S.S. Crab Nebula spectrum. The H.E.S.S. limits are considerably (>5 times) stronger than any reported to date. However, due to the generally variable nature of AGN emission, these upper limits con- strain the maximum average brightness of the AGN only during the observation time. Hence they are limits on the steady-component or quiescent flux from the AGN. Future flaring behavior may increase the VHE flux from any of these AGN to significantly higher levels. A search for VHE flux variability from each observed AGN was also performed. Here the nightly integral flux above the average energy threshold was calculated assum- ing a photon index of Γ=3.0 and fit by a constant. Any flaring behavior would be demonstrated in the form of a poor χ2 probability for the fit. Table 4 shows the dates each AGN was observed and the resulting χ2 probability. As can be seen, no evidence for VHE flux variability is found. The lack of any significant VHE detection or flaring behavior is perhaps expected from the beahvior of the in- dividual AGN in the X-ray regime. Quick-look results pro- vided by ASM/RXTE team show that none of the AGN (for which all-sky monitor data exists) were particularly active during the dates of the H.E.S.S. observations. On these dates, the measured daily average count rate from each AGN never deviated by more than ∼2σ from the mean value averaged over the whole X-ray data set. 6. Discussion Since AGN are known to emit radiation in all wave- bands, understanding and modelling their emission must take into account their entire spectral energy distribution (SED). Constraining any model is difficult as only a lim- 6 F. Aharonian et al.: Observations of Selected AGN with H.E.S.S. Table 4. The dates of the H.E.S.S. observations of each AGN and the χ2 probability for a fit of a constant to the nightly integral flux . proposed and can be constrained using the H.E.S.S. re- sults. However, some caveats are required for interpreting a blazar SED with the H.E.S.S. upper limits. MJD−50000 P(χ2) 0.85 - 0.87 0.83 0.67 - 0.85 0.13 0.72 0.90 0.50 0.71 0.81 0.35 0.25 0.13 0.55 0.52 0.89 Blazars are known to be highly variable at all wave- lengths, typically characterized by low-emission quiescent states interrupted by periods of flaring behavior where the flux increases dramatically. In some cases this in- crease is several orders of magnitude. Due to this ex- treme variability, it has been shown that fitting the SED of blazars has very large uncertainties when non- simultaneous multiwavelength (MWL) data are used (see e.g. Bottcher, Mukherjee & Reimer 2002). As a result the usefulness of non-simultaneous upper limits, as is the case for the H.E.S.S. observations, in modelling these object is limited. The H.E.S.S. upper limits, in the absence of simultaneous MWL data, are only relevant for modelling the quiescent state of the blazar using archival low-state MWL data. An additional problem using these upper lim- its arises due to the absorption of VHE photons on the EBL. The upper limit on the flux intrinsic to the object can be significantly higher than those presented in Table 3 depending on the redshift. As a result the upper limits must have the effects of the EBL removed before they can be used for modelling. Unfortunately, parameterizations of the EBL are poorly constrained leading to numerous mod- els with dramatically different behaviors, adding another significant uncertainty when using VHE upper limits to help model blazar emission. Given the wide range of EBL interpretations, this deabsorption is not performed here. Taking note of the caveats regarding the effects of the EBL and the issues with non-simultaneous observations, a comparison of the upper limits is made, where possible, to three sets of VHE flux predictions based on the SEDs of blazars. The first set (Stecker, de Jager & Salamon 1996), referred to as SDS henceforth, uses simple scaling argu- ments to predict VHE fluxes for Einstein Slew survey ob- jects. In the case of the SDS flux predictions the effects of EBL absorption are already accounted for with an "aver- aged" model. The other two sets of predictions are taken from Costamante & Ghisellini (2002). The first, referred to as FOS, uses a phenomenological description of the av- erage SED of blazars based on their bolometric luminosity (Fossati et al. 1998), modified by Donato et al. (2001), and derives predictions on the basis of the individual blazar's radio luminosity and synchrotron peak frequency. The second, referred to as CG, uses fits of a synchrotron self-Compton model to existing multiwavelength data. Both the FOS and CG predictions do not have the ef- fects of EBL absorption accounted for. This could change the flux predictions by factors of ∼5 above 300 GeV and by factors >100 above 1 TeV for objects at z∼0.2. Table 5 shows the 99.9% H.E.S.S. flux upper limits extrapolated (assuming Γ = 3.0) to above 300 GeV and above 1 TeV, as well as which predictions are available above these thresholds. The H.E.S.S. upper limits are be- low the SDS predictions above 300 GeV in three of the five cases (factors ranging from ∼2 to ∼5), and below two of the five predictions above 1 TeV (factors of ∼1.3 and Object BL Lacs 1ES 0145+138 1ES 0229+200 1ES 0323+022 PKS 0548−322 EXO 0556.4−3838 RGB J0812+026 RGB J1117+202 1ES 1440+122 RBS 1888 Q J22548−2725 PKS 2316−423 1ES 2343−151 Radio-Loud Galaxies 3202,3205,3210-14 3317 2904-05,3267 3296,3299-300,3303 3347,3354 3081 3054,3112,3114,3116 3110,3119 3207-08,3210 3201,3210-11 3201,3207-08 3211-13 3C 120 Pictor A 3C 273 Cen A 3316-19,3353-55 3269,3318-19,3351,3353-54 3109-10,3148-49 3111-13 Radio-Weak Galaxies NGC 1068 NGC 3783 NGC 7469 3290,3292-94,3296 3107-08 3202,3206,3211-12 ited number of high-energy measurements currently ex- ist (see Fichtel et al. 1994 for EGRET upper limits on blazars, Seyfert galaxies and radio-loud galaxies). This is especially true at VHE energies, making the upper lim- its presented in Table 3 quite useful due to their unprece- dented strength. While such modelling is beyond the scope of this paper, the applicability and usefulness of the limits for each of the three classes of observed AGN are dis- cussed. 6.1. BL Lacs BL Lacs belong to the sub-class of radio-loud AGN known as blazars, which are AGN thought to pos- sess a jet which is viewed close to the line of sight (Urry & Padovani 1995). The distinction between BL Lacs and other blazars is primarily based on their op- tical spectra which are characterized by weak or absent emission lines. As mentioned in the introduction, almost all VHE bright AGN belong to this class. These AGN have dominantly non-thermal emission and are character- ized by a double-humped SED. The low-energy component is widely accepted as originating from synchrotron radia- tion of relativistic electrons in the magnetic field around the object. However, the origin of the high-energy com- ponent is the subject of much debate. Various models in- volving either leptonic or hadronic processes have been F. Aharonian et al.: Observations of Selected AGN with H.E.S.S. 7 Table 5. Integral flux upper limits (99.9% confidence level, Γ=3.0) scaled to above 300 GeV, and above 1 TeV, from H.E.S.S. observations for the selected AGN where any of the SDS, FOS and CG predictions are available. The units are 10−12 cm−2 s−1. The cases for which the H.E.S.S. upper limit is below the predicted flux are shown in bold. on their radio morphology (Fanaroff & Riley 1974). FR I objects, such as Cen A (the prototype), 3C 120 and the VHE-emitter M 87, show extended jets with no distinct termination point, and FR II objects, like Pictor A, have narrow, collimated jets with terminal "hotspots." These FR objects differ from BL Lacs mainly due to a large view- ing angle (50◦−80◦) with respect to the jet axis. Blazar H.E.S.S. H.E.S.S. 300 GeV 1 TeV 0145+138 0229+200 0323+022 0548−322 0556−384 0812+026 1117+202 1440+122 2.25 5.15 1.92 2.67 7.03 4.03 5.99 4.77 0.203 0.464 0.173 0.240 0.632 0.363 0.539 0.429 Pred. SDS SDS, FOS, CG SDS, FOS, CG SDS, FOS, CG FOS FOS FOS SDS, FOS, CG ∼5). Even if the EBL absorption effects are accounted for in the SDS predictions, the discrepancies can easily be accounted for by the aforementioned simultaneity caveats and thus the H.E.S.S. upper limits do not make any strong statements regarding the SDS predictions. All the FOS predictions are above the H.E.S.S. upper limits, from fac- tors of ∼1.4 to ∼16 for the predictions above 300 GeV and factors of ∼5 to ∼40 for the predictions above 1 TeV. While at first this seems severe, accounting for TeV ab- sorption can reduce these discrepancies dramatically. In addition the FOS predictions are claimed to be more suit- able for "high" state VHE flux predictions, whereas the H.E.S.S. upper limits are most appropriate for constrain- ing the quiescent state of the AGN. Given that variability of up to two orders of magnitude have been seen in VHE blazars such as Markarian 421, it is clear that it is again difficult to test these predictions with the H.E.S.S. upper limits. However, the disagreement suggests that different sets of parameters might be necessary to account for the quiescent state of the source. The CG predictions, which are claimed to be more appropriate for the quiescent state of the AGN tested by the H.E.S.S. upper limits, are all below the upper limits. 6.2. Other Radio-Loud Galaxies Speculation exists for detectable levels of VHE emission from the jets of AGN without doppler boosting along the line of sight (see e.g. Aharonian, Coppi & Volk 1994). Therefore, the H.E.S.S. observation program also included four other radio-loud AGN. Like BL Lacs they all possess jets. One of these, 3C 273, meets some, but not all, of the phenomenological criteria for classification as a blazar. However, it is most accurately characterized as a quasar. It is the brightest and one of the most nearby quasars. The other three AGN, Cen A, 3C 120 and Pictor A, are found in Fanaroff-Riley (FR) radio galaxies. These galaxies fall into two classes, FR I and FR II. The distinction is based (for a review see Chandra observations e.g. Harris 2001) show that Pictor A, 3C 120, and 3C 273 all possess bright X-ray features like knots and hot spots in their large-scale extragalactic jets. The X-ray fluxes of these features are at least a factor of 10 larger than the radio and optical fluxes. This behavior is the opposite of the predictions from synchrotron self-Compton and inverse-Compton models and requires alternative the- oretical explanations (see e.g. Aharonian 2002). Use of the H.E.S.S. upper limits for these objects should aid in constraining some of the presented scenarios. However, they are still subject to the aforementioned variability and EBL absorption (mainly for 3C 273) caveats. Located at a distance of 3.4 Mpc, Cen A (NGC 5128) is the closest radio-loud AGN. It is one of the best-studied extragalactic objects due to its large apparent brightness in all wavebands (for a recent review see Israel 1998). The proximity of Cen A means that the intrinsic spectrum of the object is unaffected by absorption on the EBL, con- siderably simplifying the use of the H.E.S.S. upper limit in the modelling of its VHE emission. However, the lack of simultaneous observations is still an issue as Cen A, like blazars, exhibits large flux variability, albeit on much longer time scales (years). During the early days of VHE astronomy, a detection of emission above 300 GeV from Cen A was claimed using a non-imaging Cherenkov system (Grindlay et al. 1975) during a historically high emission state. The flux re- ported, I(>300 GeV) = (4.4±1.0) × 10−11 cm−2 s−1, is over an order of magnitude above the H.E.S.S. 99.9% flux upper limit extrapolated to above 300 GeV, I(>300 GeV) < 2.3 × 10−12 cm−2 s−1. The H.E.S.S. result does not contradict the claimed detection as RXTE ASM observa- tions show that Cen A was in a low emission state when observed by H.E.S.S.. During a similar low emission state, EGRET detected >100 MeV γ-ray emission from Cen A (Sreekumar et al. 1999). This is the only EGRET detec- tion associated with an AGN that is not a member of the blazar class. Extrapolating the EGRET spectrum to above the H.E.S.S. threshold yields I(>190 GeV) = 3.5 × 10−12 cm−2 s−1 which is ∼60% of the upper limit shown in Table 3. The H.E.S.S. limit is similar, 5.5 × 10−12 cm−2 s−1, when assuming the measured EGRET spectrum of Γ=2.40. These results imply that future identification of a high-emission state in Cen A should motivate further VHE observations. 8 F. Aharonian et al.: Observations of Selected AGN with H.E.S.S. 6.3. Radio-Weak Galaxies All of the radio-weak AGN observed by H.E.S.S. are lo- cated in Seyfert galaxies which differ from the galaxies pre- viously discussed in many respects. Like the other AGN, they have outflows, albeit typically with low velocity and uncollimated, approximately perpendicular to the accre- tion disk. Some even have collimated jets that emit syn- chrotron radiation. However, the jets are neither as colli- mated as in radio-loud AGN, nor do they show any indica- tions of relativistic motion. Two kinds of Seyfert galaxies (types I and II) exist whose differences can be explained in terms of viewing angle (Antonucci & Miller 1985). It is be- lieved that Seyfert I galaxies are viewed "face on" and thus the nuclear regions are directly visible, whereas Seyfert II galaxies are viewed "edge on" causing the nuclear regions to be obscured by material (the torus or warped disk). Currently, no Seyfert galaxies are known to be VHE emit- ters. Three bright well-studied Seyfert galaxies were ob- served by H.E.S.S.: NGC 1068, NGC 3783, and NGC 7469. NGC 1068 (M 77), the prototypical type II object, is the brightest and closest known Seyfert galaxy and as such is perhaps the best candidate for detection of this class at VHE energies. Here it should be noted that since the emis- sion from Seyferts is not beamed, their orientation is not as important as with blazars. NGC 3783, a classical type I ob- ject, is also one of the brightest and closest Seyfert galax- ies, and one of the most well studied. It is also interest- ing in that exceptionally deep measurements made using the Chandra X-ray Observatory reveal a fast (> 106 km hr−1) wind of highly ionized atoms blowing away from the galaxy's suspected central black hole (Kaspi et al. 2002). NGC 7469, also type I, is unusual in that it has an inner ring of gas very close to the nucleus that is undergoing massive star formation (Genzel et al. 1995). None of these objects were detected and the upper limits shown in Table 3 are quite constraining. While Seyfert-type galaxies are not necessarily expected to emit VHE γ-rays at observable levels, the H.E.S.S. results eas- ily provide constraints for modelling. This is because these AGN generally show less variability than blazars, and all the ones observed are close enough to only have mini- mal effects from the absorption of VHE photons on the EBL. The H.E.S.S. results could be interpreted as imply- ing that Seyfert-type AGN are not significant emitters of VHE photons. However, the observed sample and expo- sure times are small, making it premature to rule the class out all together. 7. Conclusions H.E.S.S. observed greater than twenty AGN in 2003 and 2004 as part of a campaign to identify new VHE-bright AGN. Several significant detections from this campaign have been reported elsewhere (see Table 1 for refer- ences). Results presented here detail the AGN observa- tions for which no significant excess was found, apart from a marginal signal from the well-known VHE-emitter Markarian 501. Despite the limited exposure (<8 hours) for each of these AGN, the upper limits on the VHE flux determined by H.E.S.S. are the most stringent to date, demonstrating the unprecedented sensitivity of the instru- ment. Clearly the strength of the limits makes them quite useful, yet it must again be stressed that any interpreta- tion using the H.E.S.S. limits must take into account both the EBL and the state of the source using simultaneous data at different wavelengths. The H.E.S.S. AGN observation program is not com- plete as many proposed candidates have not yet been ob- served. Further, more time is scheduled for observations of some of the AGN presented here as part of a monitor- ing effort for blazars. H.E.S.S. has already detected γ-ray emission from four AGN, including one (PKS 2005−489) never previously detected in the VHE regime. Clearly the prospects of finding additional VHE-bright AGN are ex- cellent. Acknowledgements. The support of the Namibian authorities and of the University of Namibia in facilitating the construc- tion and operation of H.E.S.S. is gratefully acknowledged, as is the support by the German Ministry for Education and Research (BMBF), the Max-Planck-Society, the French Ministry for Research, the CNRS-IN2P3 and the Astroparticle Interdisciplinary Programme of the CNRS, the U.K. Particle Physics and Astronomy Research Council (PPARC), the IPNP of the Charles University, the South African Department of Science and Technology and National Research Foundation, and by the University of Namibia. We appreciate the excel- lent work of the technical support staff in Berlin, Durham, Hamburg, Heidelberg, Palaiseau, Paris, Saclay, and in Namibia in the construction and operation of the equipment. References Aharonian, F.A., Coppi, P.S. & Volk, H.J., 1994, ApJ, 423, L5 Aharonian, F., Akhperjanian, A., Barrio, J., et al., 2001, A&A, 370, 112 Aharonian, F., 2002, MNRAS, 332, 215 Aharonian, F., Akhperjanian, A., Beilicke, M., et al., 2003, A&A, 403, L1 Aharonian, F., Akhperjanian, A., Beilicke, M., et al., 2004a, A&A, 421, 529 Aharonian, F., Akhperjanian, A.G., Aye ,K.-M., et al., 2004b, Astropart Phys, 22, 109 Aharonian, F., Akhperjanian, A.G., Aye ,K.-M., et al., 2005a, A&A, 430, 865 Aharonian, F., Akhperjanian, A.G., Aye ,K.-M., et al., 2005b, A&A, 436, L17 Aharonian, F., Akhperjanian, A.G., Aye ,K.-M., et al., 2005c, A&A, 437, 95 Antonucci, R.R. & Miller, J.S., 1985, ApJ, 297, 621 Beilicke, M., Cornils, R., Heinzelmann, G., et al., 2005, Proc. of the 22nd Texas Symposium on Relativistic Astrophysics (Stanford) Bernlohr, K., Carrol, O., Cornils, R., et al., 2003, Astropart Phys, 20, 111 Bottcher, M., Mukherjee, R., & Reimer, A., 2002, ApJ, 581, 143 F. Aharonian et al.: Observations of Selected AGN with H.E.S.S. 9 Bradbury, S.M., Deckers, T., Petry, D., et al., 1997, A&A, 320, L5 Buckley, J., 1999, Astropart Phys, 11, 119 Catanese, M., Akerlof, C.W., Badran, H.M., et al., 1998, ApJ, 501, 616 Chadwick, P.M., Lyons, K., McComb, T.J.L., et al., 1999a, ApJ, 513, 161 Chadwick, P.M., Lyons, K., McComb, T.J.L., et al., 1999b, ApJ, 521, 547 Chadwick, P.M., Daniel, M.K., Lyons, K., et al., 2000, A&A, 364, 450 Costamante, L. & Ghisellini G., 2002, A&A, 384, 56 de la Calle Perez, I., Bond, I.H., Boyle, P.J., et al., 2003, ApJ, 599, 909 Donato, D., Ghisellini, G., Tagliaferri, G., et al., 2001, A&A, 375, 739 Fanaroff, B.L. & Riley, J.M, 1974, MNRAS, 167, 31 Feldman, G.J. & Cousins, R.D., 1998, Phys Rev D, 57, 3873 Fichtel, C.G., Bertsch, D.L., Chiang, J., et al., 1994, ApJS, 94, 551 Fossati, G., Maraschi, L., Celotti, A., et al., 1998, MNRAS, 299, 433 Funk, S., Hermann, G., Hinton, J., et al., 2004, Astropart Phys, 22, 285 Genzel, R., Weitzel, L., Tacconi-Garman, L.E., et al., 1995, ApJ, 444, 129 Grindlay, J.E., Helmken, H.F., Brown, R.H., et al., 1975, ApJ, 197, L9 Harris, D.E, 2001, ASP Conference Proceedings, 250, 204 Hofmann, W., 2003, Proc. of the 28th ICRC (Tsukuba), 2811 Horan, D., Badran, H.M., Bond, I.H., et al., 2002, ApJ, 571, 753 Horan, D., Badran, H.M., Bond, I.H., et al., 2004, ApJ, 603, 51 Israel, F.P., 1998, A&AR, 8, 237 Kaspi, S., Brandt, W.N., George, I.M., et al., 2002, ApJ, 574, 643 Li, T. & Ma, Y., 1983, ApJ, 272, 317 Nishijima, K., 2002, Publ. Astron. Soc. Aust., 19, 26 Nishiyama, T., Chamoto, N., Chikawa, M., et al., 1999, Proc. of the 26th ICRC (Salt Lake City), 3, 370 Perlman, E.S., 1999, AIP Conf. Proc., 515, 53 Punch, M., Akerlof, C.W., Cawley, M.F., et al., 1992, Nature, 358, 477 Quinn, J., Akerlof, C.W., Biller, S.D., et al., 1996, ApJ, 456, L83 Roberts, M.D., Dazeley, S.A., Edwards, P.G., et al., 1998, A&A, 337, 25 Roberts, M.D., McGee, P., Dazeley, S.A., et al., 1999, A&A, 343, 691 Rowell, G.P., Dazeley, S.A., Edwards, P.G., et al., 1999, Astropart Phys, 11, 217 Schroedter, M., 2005, ApJ, in Press, arXiv:astro-ph/0504397 Sreekumar, P., Bertsch, D.L., Hartmann, R.C., et al., 1999 Astropart Phys, 11, 221 Stecker, F.W., de Jager, O.C., & Salamon, M.H., 1992, ApJ, 390, 49 Stecker, F.W., de Jager, O.C., & Salamon, M.H., 1996, ApJ, 473, L75 Tsvetanov, Z.I., Hartig, G.F, Ford, H.C., et al., 1998, ApJ, 493, L83 Urry, C.M. & Padovani, P., 1995, PASP, 107, 803 Vincent, P., Denance, J.-P., Huppert, J.-F., et al., 2003, Proc. of the 28th ICRC (Tsukuba), 2887 Williams, D., 2004, AIP Conference Proceedings, 745, 499
astro-ph/0408023
2
0408
2004-08-25T22:26:54
Galaxy Assembly
[ "astro-ph" ]
In a Lambda CDM Universe, galaxies grow in mass both through star formation and through addition of already-formed stars in galaxy mergers. Because of this partial decoupling of these two modes of galaxy growth, I discuss each separately in this biased and incomplete review of galaxy assembly; first giving an overview of the cosmic-averaged star formation history, and then moving on to discuss the importance of major mergers in shaping the properties of present-day massive galaxies. The cosmic-averaged star formation rate, when integrated, is in reasonable agreement with the build-up of stellar mass density. Roughly 2/3 of all stellar mass is formed during an epoch of rapid star formation prior to z=1, with the remaining 1/3 formed in the subsequent 9 Gyr during a period of rapidly-declining star formation rate. The epoch of important star formation in massive galaxies is essentially over. In contrast, a significant fraction of massive galaxies undergo a major merger at z<1, as evidenced by close pair statistics, morphologically-disturbed galaxy counts, and the build-up of stellar mass in morphologically early-type galaxies. Each of these methods is highly uncertain; yet, taken together, it is not implausible that the massive galaxy population is strongly affected by late galaxy mergers, in excellent qualitative agreement with our understanding of galaxy evolution in a Lambda CDM Universe.
astro-ph
astro-ph
Galaxy Assembly By Eric F. Bell Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany [email protected] In a ΛCDM Universe, galaxies grow in mass both through star formation and through addition of already-formed stars in galaxy mergers. Because of this partial decoupling of these two modes of galaxy growth, I discuss each separately in this biased and incomplete review of galaxy assembly; first giving an overview of the cosmic-averaged star formation history, and then moving on to discuss the importance of major mergers in shaping the properties of present-day massive galaxies. The cosmic-averaged star formation rate, when integrated, is in reasonable agreement with the build-up of stellar mass density. Roughly 2/3 of all stellar mass is formed during an epoch of rapid star formation prior to z ∼ 1, with the remaining 1/3 formed in the subsequent 9 Gyr during a period of rapidly-declining star formation rate. The epoch of important star formation in massive galaxies is essentially over. In contrast, a significant fraction of massive galaxies undergo a major merger at z < ∼ 1, as evidenced by close pair statistics, morphologically- disturbed galaxy counts, and the build-up of stellar mass in morphologically early-type galaxies. Each of these methods is highly uncertain; yet, taken together, it is not implausible that the massive galaxy population is strongly affected by late galaxy mergers, in excellent qualitative agreement with our understanding of galaxy evolution in a ΛCDM Universe. 1. Introduction The last decade has witnessed amazing progress in our empirical and theoretical un- derstanding of galaxy formation and evolution. This explosion in our understanding has been driven largely by technology: the profusion of 8 -- 10-m class telescopes, the advent of wide-field multi-object spectrographs and imagers on large telescopes, servicing missions for the Hubble Space Telescope (HST), giving it higher resolution, higher sensitivity, larger field-of-view, and access to longer wavelengths, and the commissioning and/or launch of powerful observatories in the X-ray, ultraviolet, infrared, and sub-millimeter, are but a few of the important technological advances. This has led to a much-increased empir- ical understanding of broad phenomenologies: e.g., constraints on the overall shape of the cosmic history of star formation, the increased incidence of star-forming galaxies in less dense environments, the co-evolution of stellar bulges and the supermassive black holes that they host, and the increasing incidence of galaxy interactions at progressively higher redshifts, to name but a few. In turn, tension between these new observational constraints and models of galaxy formation and evolution have spurred on increasingly complex models, giving important (and sometimes predictive!) insight into the physical processes -- star formation (SF), feedback, galaxy mergers, AGN activity -- that are driving these phenomenologies. In this review, I will give a grossly incomplete and biased overview of some hopefully interesting aspects of galaxy assembly. An important underpinning of this review is my (perhaps misguided) assumption that we live in a Universe whose broad properties are described reasonably well by the cold dark matter paradigm, with inclusion of a cosmo- logical constant (ΛCDM): Ωm = 0.3, ΩΛ = 0.7, and H0 = 70 km s−1 Mpc−1 following results from the Wilkinson Microwave Anisotropy Probe (Spergel et al. 2003) and the HST Key Project distance scale (Freedman et al. 2001). This model, while it has impor- tant and perplexing fine-tuning problems, seems to describe detection of 'cosmic jerk' using supernova type Ia (Riess et al. 2004), and the evolving clustering of the luminous 1 2 E. F. Bell: Galaxy Assembly and dark matter content of the Universe with truly impressive accuracy on a wide range of spatial scales (Seljak et al. 2004). An important feature of CDM models in general is that galaxies are formed 'bottom- up'; that is, small dark matter halos form first, and halo growth continues to the present through a combination of essentially smooth accretion and mergers of dark matter halos (e.g., Peebles 1980). Thus, small halos were formed very early, whereas larger haloes should still be growing through mergers. This has the important feature of decoupling, to a greater or lesser extent, the physics and timing of the formation of the stars in galaxies, and the assembly of the galaxies themselves from their progenitors through galaxy mergers and accretion (White & Rees 1978; White & Frenk 1991). Accordingly, in this review I strive to explore the two issues separately. First, I will explore the build-up of the stellar mass in the Universe, irrespective of how it is split up into individual galaxies. Then, I will move on to describing some of the first efforts towards understanding how the galaxies themselves assembled into their present forms. This article will not touch on many important and interesting aspects of galaxy evolution; the co-evolution (or otherwise) of galaxy bulges and their supermassive black holes (see, e.g., Peterson, these proceedings; Ferrarese et al. 2001; Haehnelt 2004), the important influences of local environment on galaxy evolution (see, e.g., Bower & Balogh 2004), or the evolution of galaxy morphology (see, e.g., Franx, these proceedings). In this review, I adopt a ΛCDM cosmology and a Kroupa (2001) IMF; adoption of a Kennicutt (1983) IMF in this article would leave the results unchanged. 2. The assembly of stellar mass In many ways, the empirical exploration of the assembly of stellar mass throughout cosmic history has been one of the defining features of the last decade of extragalactic astronomical effort. Yet, in spite of such effort, and the apparent simplicity of the goal, progress at times has been frustratingly slow. The difficulties are many: calibration of SF rate (SFR) and stellar mass indicators, the effects of dust on SFR estimates, relatively poor sensitivity for most SFR indicators, and field-to-field variations caused by large scale structure. In this section, I will briefly summarize the progress made to date towards measuring the cosmic SF history (SFH) using two independent methods: exploration of the evolution of the cosmic-averaged SFR, and the evolution of the cosmic-averaged stellar mass density, both as a function of epoch. The two are intimately related -- the cosmic SFH is the integral over the cosmic SFR -- and, barring any strong variation in stellar initial mass functions (IMFs) as a function of galaxy properties and/or epoch, consistency between the two would be expected. 2.1. The cosmic-averaged star formation rate density 2.1.1. Methodology One measures SFR by measuring galaxy luminosities in a passband or passbands which one hopes will reflect the total number of massive stars in a galaxy. Using stellar popu- lation synthesis and other models, one then attempts to convert this luminosity into a total SFR, assuming a given stellar IMF to allow conversion from the number of massive stars to the total mass in the newly-formed stellar population (see, e.g., Kennicutt 1998 for an excellent review). In some cases, this calibration from total luminosity to SFR is relatively robust, because it relies on reasonably well-understood physics. For example, the total ultraviolet (UV) light from a galaxy is reasonably robustly translated into the number of massive O and B stars. The total infrared luminosity of a starbursting galaxy, if the galaxy is E. F. Bell: Galaxy Assembly 3 optically-thick to the UV light, is a reasonable reflection of the bolometric output of this starburst. Balmer line emission, under weak assumptions about the physical conditions in Hii regions, is a reasonable indicator of the number of very massive O stars. These quantities, in turn, can be converted into a total SFR, assuming a stellar IMF which does not depend on galaxy properties and epoch. In other cases, the calibration from luminosity to SFR is much less direct: e.g., the GHz radio emission from star-forming galaxies is well-correlated with other indicators of SFR, such as IR or Balmer line luminosity, and is known to be dominated by synchrotron emission from cosmic-ray electrons spiraling in galactic magnetic fields for at least mas- sive star-forming galaxies (see Condon 1992 for an excellent review; see also Bell 2003). Yet, there is no robust theoretical understanding of why the relationship between radio emission and SFR should show such modest scatter (see, e.g., Bressan, Silva, & Granato 2002; Niklas & Beck 1997; Lisenfeld, Volk, & Xu 1996 for models of the radio emission of star-forming galaxies). On top of these interpretive challenges, there are other important difficulties. Dust extinguishes UV light very effectively; empirical dust corrections based on UV color and/or UV-optical properties calibrated on UV-bright starbursts in the local Universe (Calzetti, Kinney, & Storchi-Bergmann 1994; Calzetti 2001) do not apply to normal galaxies (Bell 2002; Kong et al. 2004), IR-bright starbursts (Goldader et al. 2002), or indeed even Hii regions (Bell et al. 2002; Gordon et al. 2004). Estimates of dust reddening from Balmer line ratios may yield reasonably accurate Balmer line-derived SFRs for ∼ 0.4. IR and galaxies (Kennicutt 1983), yet are extremely challenging to measure at z > radio facilities lack the sensitivity to probe to faint limits; only galaxies with SFRs in excess of ∼ 10 M⊙ yr−1 are observable with current facilities at z > ∼ 0.5 (e.g., Flores et al. 1999). 2.1.2. Results There are a huge number of papers which have addressed the evolution of the cosmic SFR, and are too numerous to mention or discuss in any detail. A few particularly impor- tant examples are Lilly et al. (1996), Madau et al. (1996), Flores et al. (1999), Steidel et al. (1999), and Haarsma et al. (2000). In Fig. 1, I show the general form of the cosmic SFR, as derived by Hopkins (2004) in a very nice compilation of cosmic SFR estimates, where he uses locally-calibrated relationships between dust attenuation and SFR to correct for dust. His corrections are reasonably similar to those commonly used, but with the important advantages that they are uniformly derived, and account for the well-known SFR -- dust correlation. The intention of showing this cosmic SFR is primarily to illustrate that despite the significant observational challenges and interpretive challenges it has been possible to determine its broad shape to better than a factor of three over much of cosmic history, using a variety of different observational methodologies. There are two points which one should take away from the cosmic SFR, both of which have been known at least at the qualitative level since 1996 when the first cosmic SFRs were constructed (e.g., Lilly et al. 1996; Madau et al. 1996). Firstly, it is abundantly clear that the cosmic SFR has dropped by a factor of nearly 10 since z ∼ 1.5 (see, e.g., the interesting compilation from Hogg 2001). Secondly, there is no compelling evidence against a roughly constant cosmic SFR at redshifts higher than 1 (e.g., Steidel et al. ∼ 4, 1999), with the exception of some recent explorations of UV-derived SFRs at z > which appear to be lower than those at z < ∼ 4 (e.g., Stanway et al. 2004). 4 E. F. Bell: Galaxy Assembly Figure 1. The evolution of the cosmic SFR density. SFRs assume a Kroupa (2001) IMF and H0 = 70 km s−1 Mpc−1. The data are taken from Hopkins (2004), and are corrected for dust assuming a SFR -- dust correlation as found in the local Universe. The solid line shows an empirical fit to the cosmic SFR. 2.2. The cosmic-averaged star formation history A complementary way to understand the star formation history of the Universe is to explore the build-up in stellar mass with cosmic time. This integral over the cosmic SFR contains the same information (under the assumption of a reasonably well-behaved stellar IMF), yet suffers from completely different systematic uncertainties and is therefore an invaluable probe of the broad evolution of the stellar content of the Universe. 2.2.1. Methodology Ideally stellar masses would be estimated from spectroscopic data (e.g., velocity fields, rotation curves, and/or stellar velocity dispersions) coupled with multi-waveband pho- tometry, under some set of assumptions about the dark matter content of the galaxy. Unfortunately, measurements of such quality are relatively uncommon, even in the local Universe. In the absence of velocity data, one can attempt to estimate stellar masses using photometric data alone with the aid of stellar population synthesis models (e.g., Fioc & Rocca-Volmerange 1997; Bruzual & Charlot 2003). In these models, increasing the mean stellar age or the metallicity produces almost indistinguishable effects on their broad- band optical colors, and indeed even in their spectra with the exception of a few key absorption lines (e.g., Worthey 1994). Increasing dust extinction produces very similar effects at least some of the time (e.g., Tully et al. 1998), although the relationship between reddening and extinction is rather sensitive to star/dust geometry (Witt & Gordon 2000). An increase in stellar population age, metallicity or dust content reddens and dims the stellar population, at a fixed stellar mass. Crucially, the relationship between reddening E. F. Bell: Galaxy Assembly 5 Figure 2. The evolution of the cosmic-averaged stellar mass density. Stellar masses assume a Kroupa (2001) IMF and H0 = 70 km s−1 Mpc−1. The data are taken from Rudnick et al. (2003; normalized to reproduce the z = 0 stellar mass density from Cole et al. 2001 & Bell et al. 2003), Dickinson et al. (2003), Drory et al. (2004), Glazebrook et al. (2004), and Borch et al.'s (in preparation) measurements from the COMBO-17 photometric redshift survey. The solid line shows the integral of the SFR from Fig. 1. and dimming is similar for all three effects. While this makes it extremely challenging to measure the age, metallicity or dust content of galaxies using optical broad-band colors alone, it does mean that one can invert the argument and use color and luminosity to rather robustly estimate stellar mass, almost independent of galaxy SFH, metallicity or dust content (e.g., Bell & de Jong 2001). The slope of the relationship between color and mass-to-light ratio (M/L) is passband- dependent (steeper in the blue, shallower in the near-infrared) but does not strongly depend on stellar IMF. In contrast, the zero point of the color -- M/L relation depends ∼ 1M⊙ where the bulk of strongly on stellar IMF, especially to its shape at masses < the stellar mass resides but the contribution to the total luminosity is low. There are important sources of systematic error: dust does not always move galaxies along the same color -- M/L relation as defined by age and metallicity (e.g., Witt & Gordon 2000), and most importantly, significant contributions from young stellar populations (either because the galaxy is truly young or because of recent starburst activity) can bias the stellar M/Ls at a given color towards lower values. Recently, a number of methodologies have been developed to address these limitations by inclusion of important variations in star formation history (e.g., Papovich, Dickinson, & Ferguson 2001) or by using spectral indices and colors to account for bursts and dust more explicitly (e.g., Kauffmann et al. 2003). 6 2.2.2. Results E. F. Bell: Galaxy Assembly The basic methodology has been applied in the last several years to a wide variety of galaxy surveys to explore the stellar mass density in the local Universe (e.g., Cole et al. 2001; Bell et al. 2003), its evolution out to z ∼ 1 (e.g., Brinchmann & Ellis 2000; Drory et al. 2004; Borch et al., in preparation) and even out to z ∼ 3 (e.g., Rudnick et al. 2003; Dickinson et al. 2003; Fontana et al. 2003, Glazebrook et al. 2004). In Fig. 2, I show a compilation of some of these stellar mass estimates, all transformed to a Kroupa (2001) IMF. The error bars in all cases give a rough idea of the uncertainties in stellar mass; the error bars for Drory et al. (2004), Glazebrook et al. (2004), and Borch et al. include contributions from cosmic variance also. From Figs. 1 and 2, it is clear that the epoch z > ∼ 1 is characterized by rapid star formation, with roughly 2/3 of the stellar mass in the Universe being formed in the first 5 Gyr. In contrast, from z ∼ 1 to the present day, one witnesses a striking decline in the cosmic SFR, by a factor of roughly 10. Despite this strongly suppressed SFR, roughly 1/3 of all stellar mass is formed in this interval, owing to the large amount of time between z ∼ 1 and the present day. 2.3. Are the cosmic star formation rates and histories consistent? It is interesting to test for consistency between the cosmic SFR and cosmic SFH -- a lack of consistency between these two may give important insight into the form of the stellar IMF between ∼ 1M⊙ (the mass range which optical/NIR light is most sensitive ∼ 5M⊙ (the stellar mass range probed by SFR indicators), or the calibrations of to) and > or systematic errors in stellar mass and/or SFR determinations. In order to integrate the cosmic SFR, I choose to roughly parameterize the form of the cosmic SFR following Cole et al. (2001). The cosmic SFR ψ = (0.006 + 0.072z1.35)/[1 + (z/2)2.4]M⊙yr−1 Mpc−3 provides a reasonable fit to the cosmic SFR (Fig. 1). The data are insufficient to constrain whether or not the cosmic SFR declines towards high redshifts from a maximum at z ∼ 1.5; I choose to impose a mild decrease towards high redshift, primarily because that matches the evolution in integrated stellar mass slightly better than a flat evolution. This cosmic SFR is then integrated using the PEGASE stellar population model assuming a Kroupa (2001) IMF and an initial formation redshift zf = 5. The exact integration in PEGASE accounts explicitly for the recycling of some of the initial stellar bass back into the interstellar medium; for the Kroupa (2001) IMF this fraction is ∼ 1/2, i.e., stellar mass in long-lived stars is 1/2 of the stellar mass initially formed (for a Kennicutt 1983 IMF the fraction is similar). I show the expected cosmic SFH as the solid line in Fig. 2. It is clear that the form of the cosmic SFR required in Fig. 1 reproduces rather well the cosmic SFH as presented in Fig. 2. There are some slight discrepancies; a cosmic SFR as flat as that shown in Fig. 1 appears to overpredict the amount of stellar mass that one sees at z ∼ 3. This might indicate that a drop-off in cosmic SFR towards higher redshift is required, or may indicate that an IMF richer in high-mass stars is favored for high- redshift starbursts. Yet, it is important to remember that estimates of cosmic SFR and SFH are almost impossible to nail down with better than 30% accuracy at any redshift. ∼ 1 the constraints are substantially weaker still, owing to large uncertainties from At z > large scale structure, uncertainties in the faint-end slope of the stellar mass or SFR functions used to extrapolate to total SFRs or masses, and the difficulty in measuring SFRs and masses of intensely star-forming, dusty galaxies. Therefore, I would tend to ∼ 1.5 until better and substantially deeper data are downweight this disagreement at z > available, focusing instead on the rather pleasing overall agreement between these two independent probes of the cosmic SFH. E. F. Bell: Galaxy Assembly 7 3. The assembly of galaxies §2 focused on the broadest possible picture of the assembly of stellar mass -- the build-up of the stellar population, averaged over cosmologically-significant volumes. Yet, the physical processes contributing to this evolution will be much more strongly probed by studying the demographics of the galaxy populations as they evolve. This splitting of the cosmic SFR/SFH can happen in many ways: study of the evolution of the luminosity function, stellar mass function, or SFR 'function', study of galaxies split by morphological type or rest-frame color, or by identification of galaxies during particular phases of their evolution (e.g., interactions). There has been a great deal of activity over the last decade towards this goal: implicit in the exploration of Figs. 1 and 2 is the construction of SFR and stellar mass functions, and a number of studies have explored the evolution of the galaxy population split by morphological type (e.g., Brinchmann & Ellis 2000, Im et al. 2002), rest-frame color (e.g., Lilly et al. 1996, Wolf et al. (2003), Bell et al. 2004b), or focusing on the role of galaxy interactions (e.g., Le F`evre et al. 2000, Patton et al. 2002, Conselice et al. 2003). A full and fair exploration of any or all of these goals is unfortunately beyond the scope of this work. Here, I choose to focus on one particular key issue: the importance of galaxy mergers in driving galaxy evolution in the epoch since z ∼ 3, and especially at ∼ 1. Unlike the evolution of e.g., the stellar mass function or SFR function, to which z < both quiescent evolution and galaxy accretion can contribute, galaxy mergers (especially ∼ 2) are an unmistakable hallmark of the hierarchical assembly of galaxies. those at z < Therefore, exploration of galaxy mergers directly probes one of the key features of our current cosmological model. I will focus here on exploring the number of major galaxy mergers (traditionally defined as those with mass ratios of 3:1 or less) over the last 10 Gyr. There are three complemen- tary approaches to exploring merger rate, all of which suffer from important systematic uncertainties: the evolution of the fraction of galaxies in close pairs, the evolving fraction of galaxies with gross morphological irregularities, and investigation of the evolution of plausible merger remnants. In this work, I will briefly discuss all three methods, high- lighting areas of particular uncertainty to encourage future development in this exciting and important field. 3.1. The evolution of close galaxy pairs One of the most promising measures of galaxy merger rate is the evolution in the popu- lation of close galaxy pairs. Most close (often defined as being separated by < 20 kpc), ∼ 1 Gyr owing to strong bound galaxy pairs with roughly equal masses will merge within < dynamical friction (e.g., Patton et al. 2000). Thus, if it can be measured, the fraction of galaxies in physical close pairs is an excellent proxy for merger rate. Accordingly, there have been a large number of studies which have attempted to measure close pair fraction evolution (a few examples are Zepf & Koo 1989, Carlberg, Pritchet, & Infante 1994, Le F`evre et al. 2000; see Patton et al. 2000 for an extensive discussion of the background to this subject). In Fig. 3, I show the fraction of MB − ∼ −19.5 galaxies in close ∆proj < 20h−1 kpc pairs from Patton et al. (2000), 5 log10 h < Le F`evre et al. (2000), and Patton et al. (2002). Patton et al. (2000) and Patton et al. (2002) explore the fraction of galaxies in close pairs with velocity differences < 500 kms−1, whereas Le F`evre et al. (2000) study projected galaxy pairs, and correct them for pro- jection in a statistical way. Both studies, and indeed most other studies, paint a broadly consistent picture that the frequency of galaxy interactions was much higher in the past than at the present, and that the average ∼ L∗ galaxy has suffered 0.1 -- 1 major interac- tion between z ∼ 1 and the present day. Small number statistics, coupled with differences 8 E. F. Bell: Galaxy Assembly Figure 3. The evolution of the close pair fraction. Data points show measurements of the fraction of galaxies with 1 or more neighbors within a projected distance of 20h−1 kpc, corrected for projection (open points; Le F`evre et al. 2000), and within a projected distance of 20h−1 kpc and a velocity separation of < 500 kms−1 (solid points; Patton et al. 2000, 2002). The lines show the evolution of approximately equivalent measurements from the van Kampen et al. (in prep.) mock COMBO-17 catalogs: projection-corrected fraction (dotted line), the fraction of galaxies with > 1 neighbor within 20h−1 kpc and with velocity difference < 300 kms−1 (dashed line), and the real fraction of galaxies with > 1 neighbor within 20h−1 kpc in real space. in assumptions about how to transform pair fractions into merger rate, lead to a wide dispersion in the importance of major merging since z ∼ 1. Yet, there are significant obstacles to the interpretation of these, and indeed future, insights into galaxy major merger rate evolution. Technical issues, such as contamination from foreground or background galaxies, redshift incompleteness, and the construction of equivalent samples across the whole redshift range of interest must be thought about carefully (see, e.g., Patton et al. 2000, 2002). An further concern is that most close pair measurements are based on galaxies with similar magnitudes in the rest-frame optical. Bursts of tidally-triggered star formation may enhance the optical brightness of both galaxies in pairs, and the selection of galaxies with nearly equal optical brightness may not select a sample of galaxies with nearly equal stellar masses. Bundy et al.(2004) used K-band images of a sample of 190 galaxies with redshifts to explore this source of bias, E. F. Bell: Galaxy Assembly 9 finding a substantially lower fraction of galaxies in nearly equal-mass close pairs than derived from optical data. Yet, it is also vitally important to build one's intuition as to how the measured quan- tities relate to the true quantities of interest. As a crude example of this, I carry out the simple thought experiment where we compare the pair fraction derived from projection- corrected projected pair statistics (dotted line in Fig. 3), the pair fraction derived if one used spectroscopy to keep only those galaxies with ∆v < 300 kms−1 (dashed line), and the true fraction of galaxies with real space physical separation of < 20h−1 kpc (solid line). I use a mock COMBO-17 catalog under development by van Kampen et al. (in prep.; see Van Kampen, Jimenez, & Peacock 1999 for a description of this semi-numerical tech- nique) to explore this issue; this model is used with their kind permission. It is important to note that this simulation is still under development; the match between the observa- ∼ 0.5 is mildly encouraging, although the generally flat evolution tions and models at z < of pair fraction may not be a robust prediction of this model. Nonetheless, this model can be used to great effect to gain some insight into sources of bias and uncertainty. The catalog is tailored to match the broad characteristics of the COMBO-17 survey to the largest extent possible (see Wolf et al. 2003 for a description of COMBO-17): it has 3 × 1/4 square degree fields, limited to mR < 24 and with photometric redshift accuracy mimicking COMBO-17's as closely as possible. Primary galaxies with mR < 23, 0.2 6 z 6 1.0 and MV 6 −19 are chosen; satellite galaxies are constrained only to have mR < 24 and ∆mR < 1 mag (i.e., to have a small luminosity difference). The dotted and dashed lines show the pair fraction recovered by approximately reproducing the methodologies of Le F`evre et al. (2000; dotted line) and Patton et al. (2000, 2002; dashed line). It is clear that statistical field subtraction, adopted by Le F`evre et al. (2000) and others, may be rather robust, in terms of recovering the trends that one sees with the spectroscopic+imaging data. The solid line shows the fraction of galaxies actually separated by 20h−1 kpc. It is clear that, at least in this simulation, many galaxies with projected close separation and small velocity difference are members of the primary galaxy's group which happen to lie close to the line of sight to the primary galaxy but ∼ 20 kpc from the primary. In Le F`evre et al. (2000), Patton et al. (2002), and other are > studies, this was often corrected for by multiplying the observed fraction by 0.5(1 + z) following the analysis of Yee & Ellingson (1995); the data showed in Fig. 3 was corrected in this way. Our analysis of these preliminary COMBO-17 mock catalogs suggest that these corrections are uncertain, and that our current understanding of galaxy merger rate from close pairs may be biased. It is important to remember that the simulations discussed here are preliminary; the relationship between 'observed' and true close pair fractions may well be different from the trends predicted by the model. Yet, it is nonetheless clear that further modeling work is required before one can state with confidence that one understands the implications of close pair measurements. 3.2. The evolution of morphologically-disturbed galaxies With the advent of relatively wide-area, high spatial resolution and S/N imaging from HST, it has become feasible to search for galaxy interactions by identifying morphologically- disturbed galaxies. There are a number of ways in which one can evaluate morphological disturbance: a few examples are by visual inspection (e.g., Arp 1966), residual structures in unsharp-masked or model-subtracted images (e.g., Schweizer & Seitzer 1992), auto- mated measures of galaxy asymmetry (e.g., Conselice, Bershady, & Jangren 2000), or by the distribution of pixel brightnesses and 2nd-order moment of the light distribution (e.g., Lotz, Primack, & Madau 2004). The important hallmarks that these different methodologies probe are the non-equilibrium 10 E. F. Bell: Galaxy Assembly Figure 4. The evolution of merger rate as inferred from the fraction of galaxies with gross asymmetries (Conselice et al. 2003; grey shaded region). Inferred merger rates from close pairs are reproduced from Fig. 3 for comparison. A rough fit to predictions of major merger rate (visible for 1 Gyr) from the semi-analytic galaxy formation model of Benson et al. (2002), reproduced from Conselice et al. (2003), is shown by the solid line for reference. signatures of tidal interaction -- multiple nuclei and extended tidal tails. These features are common in simulations of interacting galaxies, and cannot result from quiescent secu- lar evolution (e.g., Toomre & Toomre 1972, Barnes & Hernquist 1996). Visual inspection is a subjective way to pick out these structures, even to relatively low surface brightness limits; in contrast, automated measures of morphology typically quantify the amount of light in bright asymmetric structures, so pick out multiple bright patches or bright tidal tails rather effectively. Recently, Conselice et al. (2003) have explored the fraction of galaxies with strong asymmetries in the rest-frame B-band our to redshift ∼ 3 in the Hubble Deep Field North. Their results, for a similar absolute magnitude range as explored by Patton et al. (2002) and Le F`evre et al. (2000), are shown in Fig. 4. While small number statistics (there are less than 500 galaxies contributing to the fractions), coupled with asymmetry uncertain- ties, are clearly an issue, a broadly consistent picture is painted whereby galaxy mergers ∼ 1, and have decreased in frequency until the were substantially more frequent at z > present day. It is interesting to note that the observations are not well-reproduced by E. F. Bell: Galaxy Assembly 11 the Benson et al. or van Kampen et al. models, which both predict substantially lower merger rates. Indeed, perhaps for the first time, observers are invoking the need for many more mergers than theorists! While these first tentative steps are encouraging, there are a number of systematic uncertainties that should be considered carefully. At some level, contamination from projected close galaxy pairs is bound to be a problem. Samples of highly asymmetric galaxies will, correctly, include also some very close physically-associated galaxy pairs. Unfortunately, they will also contain a number of physically-unassociated close galaxy pairs: Le F`evre et al. (2000) find that only ∼ 30% of close galaxy pairs are likely to be physically-associated (in the sense of being in the same galaxy group; a smaller fraction still are genuine close pairs), and the mock COMBO-17 catalog analysis presented earlier suggests a fraction closer to 10%. These contaminants may be weeded out by source extraction software, as the pair of galaxy images may be parsed into 2 separate galaxies, each of which has a low asymmetry. Yet, if this image parsing is too enthusiastic, genuine major interactions may be parsed into multiple, individually rather symmetric, sub-units. Furthermore, there are differences between different classification methodologies as to what constitutes a 'merger'. An example of this is given in Fig. 18 of Conselice et al. (2004), who explore the distribution of morphologically early-type galaxies (those with dominant bulge components), late-type galaxies (those with dominant disks), and peculiar galaxies (all other types, which includes merging and irregular galaxies) in the concentration (C) and asymmetry (A) plane. C is a measure of the light concentration of a galaxy profile, having high values for centrally-concentrated light profiles, and A is a measure of asym- metry, with high values denoting asymmetric galaxies. Inspecting their Fig. 18, one sees scattered trends between visual and automated classifications. Early-type galaxies tend to have higher concentrations and lower asymmetries, later types have lower concentra- tions and lower asymmetries, and peculiar systems have low concentrations and a wide range of asymmetries, with a tail to very high asymmetries. Yet, there are a number of early-types with low concentrations, and importantly a large number of peculiar galaxies have low asymmetries, and cannot be distinguished from morphologically-normal galax- ies in this scheme. Preliminary comparisons between by-eye morphologies and C and A values for F606W images of galaxies in the GEMS (Galaxy Evolution from Morphol- ogy and SEDs; Rix et al. 2004) HST survey supports this conclusion, indicating that such automated schemes find it hard to differentiate between irregular galaxies (whose morphologies, to the human eye, indicate sporadic star formation triggered by internal processes) and clearly interacting galaxies with morphological features indicative of tidal interactions, such as multiple bright nuclei or tidal tails. A further complication is that the human eye is more sensitive to faint tidal tails in early- or late-stage mergers than automated schemes (i.e., the timescale probed by visual classifications may be longer than for automated schemes). This discussion is by no means meant to detract from the value of the intriguing and ground-breaking work on merger rates from morphology; rather it is to emphasize that the measurement of merger rate is a subtle and difficult endeavor, fraught with systematic uncertainties which will likely have to be modeled explicitly using future generations of N -body, semi-analytic and SPH simulations, coupled with the artificial redshifting experiments already commonly used in this type of work. 3.3. The evolution of plausible merger remnants From the earliest days of galaxy morphological classification, a population of galaxies whose light distribution is dominated by a smoothly distributed, spheroidal, centrally- concentrated light distribution was noticed. These early-type galaxies are largely sup- 12 E. F. Bell: Galaxy Assembly ported by the random motions of their stars (Davies et al. 1983). These properties are very naturally interpreted as being the result of violent relaxation in a rapidly-changing potential well (Eggen, Lynden-Bell, & Sandage 1962). Therefore, in our present cosmo- logical context, these galaxies are readily identified with the remnants of major galaxy mergers (e.g., Toomre & Toomre 1972; Barnes & Hernquist 1996). Detailed comparisons of simulated major merger remnants broadly supports this notion although some inter- esting discrepancies with observations remain, and are perhaps telling us about difficult- to-model gas-dynamical dissipative processes (e.g., Bendo & Barnes 2000; Meza et al. 2003; Naab & Brukert 2003). There is strong observational support for this notion -- the correlation between fine morphological structure and residuals from the color -- magnitude correlation (Schweizer & Seitzer 1992), the existence of kinematically-decoupled cores (e.g., Bender 1988), and the similarity between the stellar dynamical properties of late- stage IR-luminous galaxy mergers and elliptical galaxies (e.g., Genzel et al. 2001). Therefore, study of the evolution of spheroid-dominated, early-type galaxies may be able to give important insight into the importance of galaxy merging through cosmic history. There are, as always, a variety of important complications and limitations to this approach. For example, a fraction of the galaxies becoming early-type during galaxy mergers will later re-accrete a gas disk, which will gradually transform into stars, making the galaxy into a later-type galaxy with a substantial bulge component (e.g., Baugh, Cole, & Frenk 1996). Furthermore, not all galaxy mergers will result in a spheroid- dominated galaxy; some lower mass-ratio interactions will result in a disk-dominated galaxy (e.g., Naab & Burkert 2003). In addition, it would be foolish to a priori ignore the possibility that an important fraction of early-type galaxies may be formed rapidly in mergers of very gas-rich progenitors at early epochs -- a scenario reminiscent of the classical monolithic collapse picture (Larson 1974; Arimoto & Yoshii 1987). Yet, these interpretive complications, much as in all the cases discussed above, do not lessen the value of placing observational constraints on the phenomenology, with the confidence that our interpretation and understanding of the phenomenology will improve with time. The rate of progress in this field has largely been determined by the availability of wide-format high-resolution imaging from HST. Ground-based resolution is insufficient to robustly distinguish disk-dominated and spheroid-dominated galaxies at cosmologically- interesting redshifts. In the local Universe, the vast majority of morphologically early- type galaxies occupy a relatively tight locus in color -- magnitude space -- the color -- magnitude relation (e.g., Sandage & Visvanathan 1978; Bower, Lucey, & Ellis 1992; Schweizer & Seitzer 1992; Strateva et al. 2001). Therefore, many workers have focused on the evolution of the red galaxy population as an accessible alternative. The efficacy of this approach is only recently being tested. Accordingly, I explore the evolution of the red galaxy population first, turning subsequently to the evolution of the early-type population later. 3.3.1. The evolution of the total stellar mass in red-sequence galaxies It has become apparent only in the last 5 years that the color distribution of galaxies is bimodal in both the local Universe (e.g., Strateva et al. 2001; Baldry et al. 2004) and out to z ∼ 1 (Bell et al. 2004b). This permits a model-independent definition of red galaxies -- those that lie on the color -- magnitude relation. A slight complication is that the color of the red sequence evolves with time as the stars in red-sequence galaxies age, necessitating an evolving cut between the red sequence and the 'blue cloud'. The evolution of this cut means that some workers who explore the evolution of red galaxies using rather stringent color criteria -- e.g., galaxies the color of local E/S0 galaxies E. F. Bell: Galaxy Assembly 13 Figure 5. The evolution of the stellar mass density in red-sequence galaxies. Stellar masses assume a Kroupa (2001) IMF and H0 = 70 km s−1 Mpc−1. The local data point is taken from Bell (2003), and data for 0.2 < z < 1.3 is taken from Bell et al. (2004b), Chen et al. (2003), and Cimatti et al. (2002). The solid line shows a rough fit to the total stellar mass density in red-sequence galaxies in the Cole et al. (2000) semi-analytic galaxy formation model. The dotted line shows the expected result if red-sequence galaxies were formed at z ≫ 1.5 and simply aged to the present day. -- find much faster evolution in the red galaxy population than those who adopt less stringent or evolving cuts (e.g., Wolf et al. 2003). Bearing this in mind, I show the evolution in the total stellar mass in red galaxies in Fig. 5. Although many surveys could in principle address this question (e.g., the surveys discussed in §2.2), very few have split their stellar mass evolution by color, and to date most surveys have only published evolution of luminosity densities in red galaxies. The z = 0 stellar mass density in red galaxies is from SDSS and 2MASS (Bell et al. 2003). The COMBO-17 datapoints for jB evolution (Bell et al. 2004b) are converted to stellar mass using color-dependent stellar M/Ls as defined by Bell & de Jong (2001); extrapolation to total stellar mass density adopts a faint-end slope α = −0.6, as found ∼ 1. Error bars account by Bell et al. (2003; 2004b) for red-sequence galaxies at z < for stellar mass uncertainties and cosmic variance, defined by the field-to-field scatter in stellar mass densities from the 3 COMBO-17 fields. Stellar masses for the LCIRS sample of red galaxies were estimated using the rest-frame R-band luminosity density presented by Chen et al. (2003), accounting for the mildest possible passive luminosity evolution (to account for evolution in stellar population color and luminosity in the most conservative way, so as to minimize any stellar mass evolution), and using a stellar 14 E. F. Bell: Galaxy Assembly M/L for early-type galaxies from Bell & de Jong (2001) using a Kroupa (2001) IMF, and adopting a color of B − R = 1.5 for early-type galaxies as a z = 0 baseline. Error bars for include stellar mass estimation uncertainties and estimated cosmic variance, following Somerville et al. (2004). The K20 data point at z ∼ 1.1, derived from ERO 'old' galaxy space densities from Cimatti et al. (2002), is very roughly calculated using a number of doubtless poor assumptions. A (rather large) stellar mass of ∼ 1011M⊙ is attached to each galaxy, and the densities are multiplied by 2 to account for the star-forming EROs (the split was roughly 50:50). This stellar mass density was multiplied by 2 again to account for fainter, undetected galaxies. Error bars of ±0.2 dex and ±0.3 dex, combined in quadrature, account for cosmic variance following Somerville et al. (2004) plus our poor modeling assumptions. Owing to their use of discordant cosmologies, I do not show the inferred stellar mass evolution of the CFRS red galaxies in Fig. 5; however, like Bell et al. (2004b) they infer no evolution in the rest-frame B-band luminosity density to within their errors (Lilly et al. 1995), meaning that their stellar mass evolution would fall into excellent agreement with those of Bell et al. (2004b) or Chen et al. (2003). The results are shown in Fig. 5. To first order, the luminosity density in the optical in red galaxies is constant out to z ∼ 1; this is confirmed by a number of surveys (e.g., Lilly et al. 1995, Chen et al. 2003, or Bell et al. 2004b). Coupled with the passive ageing of the stellar populations of these red galaxies (as is indicated by their steady reddening with cosmic time, and is confirmed by study of dynamically-derived M/Ls and absorption line ratios; e.g., Wuyts et al. 2004; Kelson et al. 2001), this implies a steadily increasing stellar mass density in red galaxies since z ∼ 1.2. To date, to the best of our knowledge, there are no published determinations of red galaxy stellar mass or luminosity density which contradict this picture. The implications of this result are rather far-reaching. Bearing in mind that at z > ∼ 1 that the red galaxy population may be significantly contaminated and/or dominated by dusty star-forming galaxies (e.g., Yan & Thompson 2003, Moustakas et al. 2004), this evolution may well represent a strong upper limit to the stellar mass in early-type galaxies since z ∼ 1.2, unless large populations of very bright blue morphologically early-type galaxies are found. I address this question next, by exploring the stellar mass evolution in morphologically early-type galaxies since z ∼ 1. 3.3.2. The evolution of the total stellar mass in early-type galaxies Owing to the lack of extensive wide-area HST-resolution imaging data, there are even weaker constraints on the evolution of stellar mass in morphologically early-type galaxies. We show published results from Im et al.(2002), Conselice et al. (2004), and Cross et al. (2004), supplementing them with a preliminary analysis of galaxies from GEMS and COMBO-17, which is presented with the permission of the GEMS and COMBO-17 teams. Im et al.(2002) present a thorough study of the luminosity density evolution of E/S0 galaxies from the DEEP Groth Strip Survey, defined using bulge -- disk decompositions and placing a limit on residual substructure in the model-subtracted images, supplemented with HST V and I-band imaging for 118 square arcminutes. A final sample of 145 E/S0 galaxies with 0.05 < z < 1.2 are isolated. The fit of average early-type galaxy stellar M/L as a function of redshift from COMBO-17/GEMS; log10 M/LB = 0.34 − 0.27z, was used to transform the published B-band luminosity densities into stellar mass densi- ties for the purposes of Fig. 6†. Cross et al. (2004) present rest-frame B-band luminosity functions for visually classified E/S0 galaxies in 5 ACS fields: the B-band luminosity den- † Again, a Kroupa (2001) IMF was used. E. F. Bell: Galaxy Assembly 15 Figure 6. The evolution of the stellar mass density in early-type galaxies. Stellar masses assume a Kroupa (2001) IMF and H0 = 70 km s−1 Mpc−1. The left-hand panel shows the stellar mass densities from GEMS only; solid circles denote early-type galaxies and open circles red-sequence galaxies. The ratio of stellar mass in early-type galaxies to the red sequence is shown in the lower left panel; the asterisk is the local value taken from Bell (2003), and the grey line shows a linear fit to the GEMS data only (RMS ∼ 15%). The right-hand panel shows the resulting corrected early-type galaxy stellar mass density evolution, where again the local data are taken from Bell (2003), the solid circles denote the GEMS+COMBO-17 result, the diamonds show results from Im et al.(2002), the naked error bars show results from Conselice et al. (2004), and the asterisks show results from Cross et al. (2004). sities were transformed into stellar mass in exactly the same way as for Im et al.(2002). Conselice et al. (2004) presented stellar mass densities directly, and these are reproduced on Fig. 6, after accounting for our use of a Kroupa (2001) IMF. The preliminary GEMS/COMBO-17 early-type galaxy data points depict the evolution of total stellar mass in galaxies with S´ersic indices n > 2.5. In GEMS, Bell et al. (2004a) found that galaxies with n > 2.5 included ∼ 80% of the visually-classified E/S0/Sa galaxy population at z ∼ 0.7, with ∼ 20% contamination from later galaxy types (recall, high S´ersic indices indicate concentrated light profiles). Here, in this very preliminary exploration of the issue, a n = 2.5 cut is adopted irrespective of redshift, ignoring the important issue of morphological k-correction. To recap, stellar masses are calculated using color-dependent stellar M/Ls from Bell & de Jong (2001), again extrapolating to total using a faint-end slope α = −0.6. The resulting n > 2.5 stellar mass evolution for the 30′ × 30′ GEMS field is shown in the left panel of Fig. 6. One can clearly see strong variation in the stellar mass density of early-type galaxies, resulting from large scale structure along the line of sight. From such data, it is clearly not possible to place any but the most rudimentary and uninteresting limits on the evolution of early-type galaxies over the last 9 Gyr. Yet, noting that the bulk of early-type galaxies are in the red sequence at z ∼ 0 (e.g., Strateva et al. 2001) and at z ∼ 0.7 (e.g., Bell et al. 2004a), and that the stellar mass density of red-sequence galaxies undergo very similar fluctuations, it becomes interesting to ask if the ratio of stellar mass in red-sequence galaxies to early types varies more smoothly with redshift. 16 E. F. Bell: Galaxy Assembly This is plotted in the lower left panel of Fig. 6. There is a weak trend in early-type galaxy to red-sequence galaxy mass density, caused by an increasingly important population of blue galaxies with n > 2.5 towards higher and higher redshift (see also Cross et al. 2004, who discuss this issue in much more depth). Importantly, however, there is only a ∼ 15% scatter around this trend despite the nearly order of magnitude variation in galaxy density, arguing against strong environmental dependence in the early type to red galaxy ratio. This relatively slow variation in early-type to red-sequence ratio (modeled using a simple linear fit for the purposes of this paper) is used to convert COMBO-17's stellar mass evolution in red galaxies from Fig. 5, which is much less sensitive to large scale structure, into the evolution of stellar mass density in morphologically early-type galaxies, which is shown in the right panel of Fig. 6. It is clear that there is a strongly increasing stellar mass density in morphologically early-type galaxies since z ∼ 1.2†. While there are ∼ 0.8 recent indications that lower-luminosity early-type galaxies are largely absent at z > (e.g., Kodama et al. 2004, de Lucia et al. 2004), the total stellar mass density is strongly dominated by ∼ L∗ galaxies, and there is no room to avoid the conclusion that there has been a substantial build-up in the total number of ∼ L∗ early-type galaxies since z ∼ 1.2. Like the results presented in sections 3.1 and 3.2, these results suggest an important role for z < ∼ 1 galaxy mergers in shaping the present-day galaxy population‡. 3.4. The importance of galaxy mergers since z ∼ 1 Three independent methods have been brought to bear on this problem -- the evolution of close galaxy pairs, the evolution of morphologically-disturbed galaxies, and the evolu- tion of early-type galaxies as a proxy for plausible merger remnants. All three methods suffer from important systematic and interpretive difficulties. Close galaxy pairs, even ∼ 500 kms−1 of each other, supplemented with spectroscopy to isolate galaxies within < remain contaminated by galaxies in the same local structures. Some measures of mor- phological disturbance are susceptible to this source of error to a lesser extent, and a clear consensus on the meaning of the different automated and visual measures of mor- phological disturbance is yet to emerge. Early-type galaxies will be the result of only a subset of galaxy mergers and interactions, and the role of disk re-accretion and fading in driving early-type galaxy evolution is frustratingly unclear. Yet, despite these difficulties some broad features are clear. Mergers between ∼ L∗ galaxies are almost certainly more frequent at higher redshift than at the present day, but this does not imply by any means that galaxy mergers are unimportant at z ∼ 1: indeed, it is possible that an average ∼ L∗ galaxy undergoes roughly 1 merger between z = 1 and the present day (Le F`evre et al. 2000). This is supported by the substantial build-up in stellar mass in red-sequence and early-type galaxies since z ∼ 1. Substantial uncertainties remain and important questions, like the stellar mass dependence of the † Im et al.(2002) report a low stellar mass density in early-type galaxies at 0.05 < z < 0.6. We would attribute this deficiency in stellar mass to incompleteness at bright apparent magnitudes, leading to a deficit of nearby E/S0s with large stellar masses (as argued by Im et al. themselves), and perhaps to small number statistics and cosmic variance (as the volume probed by this study at z < ∼ 0.5 is rather small). Further work is required to explore further this discrepancy. ‡ Disk re-accretion and fading work in opposite directions in this context; disk accretion fol- lowing a major merger takes galaxies away from the early-type class, whereas fading of disks which formed at earlier times will increase the relative prominence of spheroidal bulge compo- nents and add galaxies to the early-type class. A thorough investigation of these issues, involving bulge -- disk decompositions of rest-frame B-band images of the GEMS sample and drawing on galaxy evolution models to build intuition of the importance and effects of different physical processes, is in preparation (Haussler et al., in prep.). E. F. Bell: Galaxy Assembly 17 merger rate, or the fraction of dissipationless vs. dissipational mergers, are completely open. Nonetheless, these first, encouraging steps imply that the massive galaxy popula- tion is strongly affected by late galaxy mergers, in excellent qualitative agreement with our understanding of galaxy evolution in a ΛCDM Universe. 4. Summary and Outlook The last decade has witnessed impressive progress in our understanding of galaxy formation and evolution. Despite important technical and interpretive difficulties, the broad phenomenology of galaxy assembly has been described with sufficient accuracy to start constraining models of galaxy formation and evolution. The history of star formation has attacked from two complementary and largely independent angles -- through the evolution of the cosmic-averaged star formation rate, and through the build-up of stellar mass with cosmic time. Prior to z ∼ 1, stars formed rapidly, and ∼ 2/3 of the total present-day stellar mass was formed in this short ∼ 4 Gyr interval. The epoch subsequent to z ∼ 1 has seen a dramatic decline in cosmic-averaged SFR by a factor of roughly 10; however, ∼ 1/3 of the present-day stellar mass was formed in this 9 Gyr interval. These two diagnostics of cosmic star formation history paint a largely consistent picture, giving confidence in its basic features. The assembly of present-day galaxies from their progenitors through the process of galaxy mergers was studied using three largely independent diagnostics -- the evolu- tion of galaxy close pairs, the evolution of morphologically-disturbed galaxies, and the evolution of early-type galaxies as a plausible major merger remnant. The interpretive difficulties plaguing all three diagnostics are acute; accordingly our understanding of the importance of major mergers in shaping the present-day galaxy population is in- complete. Yet, all three diagnostics seem to indicate that an important fraction (dare I ∼ 1/2?) of ∼ L∗ ∼ 3 × 1010M⊙ galaxies are affected by galaxy mergers since suggest > z ∼ 1. This contrasts sharply with the form of the cosmic star formation history, where all of the action is essentially over by z ∼ 1. It is not by any means indefensible to argue that late mergers shape the properties of the massive galaxy population in a way which is qualitatively consistent with our understanding of galaxy formation and evolution in a ΛCDM Universe. Yet, it is clear that much work is to be done to fully characterize the physical processes ∼ 6. The 'shopping list' is too driving galaxy evolution in the epoch since reionization z < extensive to discuss properly, so I will focus on three aspects which I feel are particularly important. 4.1. Our increasing understanding of the local Universe Wide area, uniform photometric and spectroscopic surveys, such as the Sloan Digital Sky Survey (SDSS; York et al. 2000), the Two Micron All Sky Survey (2MASS; Skrutskie et al. 1997), and the Two degree Field Galaxy Redsift Survey (2dFGRS; Colless et al. 2001) are revolutionizing our understanding of the local galaxy population. These surveys have allowed one to tie down the z = 0 data point for many evolutionary studies, greatly increasing the redshift and time-baseline leverage. Yet, this increased leverage is often difficult to fully apply, because local studies suffer from very different systematics than the lookback studies, and experimental details such as imaging depth, resolution, waveband, etc. are often imperfectly matched. At this stage, few groups have grasped the nettle of repackaging these local surveys in a format which can be readily artificially redshifted, etc. to allow for uniform analysis of the distant and local control samples. This will be an important feature of the most robust of the future works in galaxy evolution, and will 18 E. F. Bell: Galaxy Assembly greatly increase the scope and discriminatory power of studies of galaxy photometric, dynamical, and morphological evolution. 4.2. Using models as interpretive tools Large and uniform multi-wavelength and/or spectroscopic datasets are becoming increas- ingly common. An important consequence of this change in the nature of the data used to study galaxy evolution is that often the error budget is completely dominated by systematic and interpretive uncertainties. In this context, models of galaxy formation and evolution can help to understand and limit these systematic uncertainties through exhaustive analysis of realistic mock catalogs. A crude example of this kind of analysis is given in §3.1; a very nice example is the 2dFGRS group catalog of Eke et al. (2004). Mock simulations of this kind of quality must become a more prominent part of our toolkit in order for the kind of interpretive difficulties faced in §3.1 or §3.3 to be successfully navigated. 4.3. The importance of multiple large HST fields Large scale structure is an important and frequently neglected source of systematic un- certainty (Somerville et al. 2004)†. Comparison of the left-hand and right-hand panels of Fig. 6 illustrates this point powerfully; the broad features of the evolution of early-type galaxy stellar mass density are discernable using the 30′ × 30′ Chandra Deep Field South alone, but correction of this result for cosmic variance using the other two COMBO-17 fields yields a significantly more convincing picture. Yet, while this kind of correction for cosmic variance may work when there is significant overlap between populations of interest (although it is debatable how far one should push such an idea), it is not a priori clear that rarer and/or optically-obscured phases of galaxy evolution such as AGN or IR-luminous mergers will be well modelled with such techniques. Furthermore, short-timescale astronomical phenomena, such as AGN or galaxy mergers, have lower number density and are potentially very strongly clustered leading to large uncertainties from number statistics and cosmic variance. Yet, these phases of galaxy evolution, where galaxies undergo important and potentially permanent transformations in the cosmic blink of an eye, require HST-resolution data in order to explore their physical drivers. When HST could be viewed as an essentially endless resource, a piecemeal approach was perfectly optimal: more and/or larger HST fields could be justified on a case-by-case basis, depending on the science goals of interest. Yet, faced with an unclear future for HST, it is not obvious that this approach is optimal. HST perhaps should be thought of as a fixed lifetime experiment, where the primary goal could become the creation of an archival dataset which will support 10 -- 20 years of top-class research. In the creation of such an archival dataset, important questions will need to be addressed: availabil- ity of resources will need to be balanced against number statistics and cosmic variance, arguably at least 2 HST passbands will be required to allow some attempt at morpholog- ical k-correction, and deep multi-wavelength data will be required for study of black hole accretion and obscured star formation, naturally driving the fields into one of a small number of low HI and cirrus holes (see Papovich, these proceedings). I wish to warmly thank the GEMS and COMBO-17 collaborations -- Marco Barden, Steven Beckwith, Andrea Borch, John Caldwell, Simon Dye, Boris Haussler, Catherine Heymans, Knud Jahnke, Martina Kleinheinrich, Shardha Jogee, Daniel McIntosh, Klaus † It is interesting to note that a ×10 increase in area in a single contiguous field gives only a ×2 reduction in cosmic variance, because the various parts of the single field are correlated with each other. E. F. Bell: Galaxy Assembly 19 Meisenheimer, Chien Peng, Hans-Walter Rix, Sebastian Sanchez, Rachel Somerville, Lutz Wisotzki, and last but by no means least Christian Wolf -- for their permission to present some GEMS and COMBO-17 results before their publication, for useful discussions, and for their friendship and collaboration. It is a joy to be part of these teams. I wish also to thank Eelco van Kampen and his collaborators for their efforts to construct mock COMBO-17 catalogs, for their permission to share results from these catalogs in this article, and for useful comments. Chris Conselice, Emmanuele Daddi, Sadegh Kochfar, and Casey Papovich are thanked for useful and thought-provoking discussions on some of the topics discussed in this review, and Chris Conselice and Shardha Jogee are thanked for constructive comments on the first version of this review. This work is supported by the European Community's Human Potential Program under contract HPRN-CT-2002- 00316, SISCO. REFERENCES Arimoto, N., & Yoshii, Y. 1987 Chemical and photometric properties of a galactic wind model for elliptical galaxies A&A 173, 23 -- 38. Arp, H. 1966 Atlas of Peculiar Galaxies ApJS 14, 1 -- 20. Baldry, I. K., Glazebrook, K., Brinkmann, J., Ivezic, Z., Lupton, R. H., Nichol, R. C., & Szalay, A. S. 2004 Quantifying the Bimodal Color -- Magnitude Distribution of Galaxies ApJ 600, 681 -- 694. Barnes, J. E., Hernquist, L. 1996 Transformations of Galaxies II. Gasdynamics in Merging Disk Galaxies ApJ 471, 115 -- 142. Baugh, C. M., Cole, S., & Frenk, C. S. 1996 Evolution of the Hubble sequence in hierarchical models for galaxy formation MNRAS 283, 1361 -- 1378. Bell, E. F. 2002 Dust-induced Systematic Errors in Ultraviolet-derived Star Formation Rates ApJ 577, 150 -- 154. Bell, E. F. 2003 Estimating Star Formation Rates from Far-infrared and Radio Luminosities: The Origin of the Radio -- Far-infrared Correlation ApJ 586, 794 -- 813. Bell, E. F., & de Jong, R. S. 2001 Stellar Mass-to-Light Ratios and the Tully -- Fisher Relation ApJ 550, 212 -- 229. Bell, E. F., Gordon, K. D., Kennicutt, Jr., R. C., & Zaritsky, D. 2002 The Effects of Dust in Simple Environments: Large Magellanic Cloud Hii Regions ApJ 565, 994 -- 1010. Bell, E. F., McIntosh, D. H., Katz, N., & Weinberg, M. D. 2003 The Optical and Near-IR Properties of Galaxies: I. Luminosity and Stellar Mass Functions ApJS 149, 289 -- 312. Bell, E. F., et al. 2004a GEMS Imaging of Red Sequence Galaxies at z ∼ 0.7: Dusty or Old? ApJ 600, L11 -- 14. Bell, E. F., et al. 2004b Nearly 5000 Distant Early-Type Galaxies in COMBO-17: A Red Sequence and Its Evolution since z ∼ 1 ApJ 608, 752 -- 767. Bender, R. 1988 Rotating and counter-rotating cores in elliptical galaxies A&A 202, L5 -- 8. Bendo, G. J., & Barnes, J. E. 2000 The line-of-sight velocity distributions of simulated merger remnants MNRAS 316, 315 -- 325. Benson, A. J., Lacey, C. G., Baugh, C. M., Cole, S., & Frenk, C. S. 2002 The effects of photoionization on galaxy formation -- I. Model and results at z = 0 MNRAS 333, 156 -- 176. Bower, R. G., & Balogh, M. L. 2004 The Difference Between Clusters and Groups: A Journey from Cluster Cores to Their Outskirts and Beyond Clusters of Galaxies: Probes of Cosmological Structure and Galaxy Evolution (eds. J. S. Mulchaey, A. Dressler & A. Oemler), p. 326 Cambridge University Press Bower, R. G., Lucey, J. R., & Ellis, R. S. 1992 Precision Photometry of Early Type Galax- ies in the Coma and Virgo Clusters: a Test of the Universality of the Colour -- magnitude Relation -- II. Analysis MNRAS 254, 601 -- 603. Brinchmann, J., & Ellis, R. S. 2000 The Mass Assembly and Star Formation Characteristics of Field Galaxies of Known Morphology ApJ 536, L77 -- 80. 20 E. F. Bell: Galaxy Assembly Bressan, A., Silva, L., & Granato, G. L. 2002 Far infrared and radio emission in dusty starburst galaxies A&A 392, 377 -- 391. Bruzual, G., & Charlot, S. 2003 Stellar population synthesis at the resolution of 2003 MNRAS 344, 1000 -- 1028. Bundy, K., Fukugita, M., Ellis, R. S., Kodama, T., & Conselice, C. J. 2004 A Slow Merger History of Field Galaxies since z ∼ 1 ApJ 601, L123 -- 126. Calzetti, D. 2001 The Dust Opacity of Star-forming Galaxies PASP 113, 1449 -- 1485. Calzetti, D., Kinney, A. L., & Storchi-Bergmann, T. 1994 Dust extinction of the stellar continua in starburst galaxies: The ultraviolet and optical extinction law ApJ 429, 582 -- 601. Carlberg, R. G., Pritchet, C. J., & Infante, L. 1994 A survey of faint galaxy pairs ApJ 435, 540 -- 547. Chen, H.-W., et al. 2003 The Las Campanas Infrared Survey IV. The Photometric Redshift Survey and the Rest-Frame R-Band Galaxy Luminosity Function at 0.5 6 z 6 1.5 ApJ 586, 745 -- 764. Cimatti, A., et al. 2002 The K20 survey -- I. Disentangling old and dusty star-forming galaxies in the ERO population A&A 381, L68 -- 72. Cole, S., Lacey, C. G., Baugh, C. M., Frenk, C. S. 2000 Hierarchical galaxy formation MNRAS 319, 168 -- 204. Cole, S., et al. 2001 The 2dF galaxy redshift survey: near-infrared galaxy luminosity functions MNRAS 326, 255-273. Colless, M., et al. 2001 The 2dF Galaxy Redshift Survey: spectra and redshifts MNRAS 328, 1039 -- 1063. Condon, J. J. 1992 Radio emission from normal galaxies ARA&A 30, 575 -- 611. Conselice, C. J., Bershady, M. A., Jangren, A. 2000 The Asymmetry of Galaxies: Physical Morphology for Nearby and High-Redshift Galaxies ApJ 529, 886 -- 910. Conselice, C. J., Bershady, M. A., Dickinson, M., & Papovich, C. 2003 A Direct Mea- surement of Major Galaxy Mergers at z < ∼ 3 AJ 126, 1183 -- 1207. Conselice, C. J., Blackburne, J. A., & Papovich, C. 2004 The luminosity, stellar mass, and number density evolution of field galaxies of known morphology from z = 0.5 − 3 submitted to the ApJ (astro-ph/0405001). Cross, N. J. C., et al. 2004 The luminosity function of early-type field galaxies at z ∼ 0.75 AJ, in press (astro-ph/0407644). Davies, R. L., Efstathiou, G., Fall, S. M., Illingworth, G., Schechter, P. L. 1983 The kinematic properties of faint elliptical galaxies ApJ 266, 41 -- 57. de Lucia, G., et al. 2004 The Buildup of the Red Sequence in Galaxy Clusters since z ∼ 0.8 ApJ 610, L77 -- 80. Dickinson, M. E., Papovich, C., Ferguson, H. C., & Budav´ari, T. 2003 The Evolution of the Global Stellar Mass Density at 0 < z < 3 ApJ 587, 25 -- 40. Drory, N., Bender, R., Feulner, G., Hopp, U., Maraston, C. Snigula, J., & Hill, G. J. 2004 The Munich Near-Infrared Cluster Survey (MUNICS) VI. The Stellar Masses of K-Band-selected Field Galaxies to z ∼ 1.2 ApJ 608, 742 -- 751. Eggen, O. J., Lynden-Bell, D., Sandage, A. R. 1962 Evidence from the motions of old stars that the Galaxy collapsed ApJ 136, 748 -- 766. Eke, V. R., et al. 2004 Galaxy groups in the 2dFGRS: the group-finding algorithm and the 2PIGG catalogue MNRAS 348, 866 -- 878. Ferrarese, L., Pogge, R. W., Peterson, B. M., Merritt, D., Wandel, A., Joseph, C. L. 2001 Supermassive Black Holes in Active Galactic Nuclei -- I. The Consistency of Black Hole Masses in Quiescent and Active Galaxies ApJ 555, L79 -- 82. Fioc, M., & Rocca-Volmerange, B. 1997 PEGASE: a UV to NIR spectral evolution model of galaxies. Application to the calibration of bright galaxy counts. A&A 326, 950 -- 962. Flores, H., et al. 1999 15µm ISO Observations of the 1415+52 CFRS Field: The Cosmic SFR as Derived from Deep UV, Optical, Mid-IR, and Radio Photometry ApJ 517, 148 -- 167. Fontana, A., et al. 2003 The Assembly of Massive Galaxies from Near-Infrared Observations of the Hubble Deep Field South ApJ 594, L9 -- 12. Freedman, W. L., et al. 2001 Final Results from the Hubble Space Telescope Key Project to Measure the Hubble Constant ApJ 553, 47 -- 72. E. F. Bell: Galaxy Assembly 21 Genzel, R., Tacconi, L. J., Rigopoulou, D., Lutz, D., Tecza, M. 2001 Ultraluminous Infrared Mergers: Elliptical Galaxies in Formation? ApJ 563, 527 -- 545. Glazebrook, K., et al. 2004 A high abundance of massive galaxies 3 -- 6 billion years after the Big Bang Nature 430, 181 -- 184. Goldader, J. D., et al. 2002 Far-Infrared Galaxies in the Far-Ultraviolet ApJ 568, 651 -- 678. Gordon, K. D., et al. 2004 Spatially Resolved Ultraviolet, Hα, Infrared, and Radio Star Formation in M81 ApJS in press. (astro-ph/0406064) Haarsma, D. B., Partridge, R. B., Windhorst, R. A., & Richards, E. A. 2000 Faint Radio Sources and Star Formation History ApJ 544, 641 -- 658. Haehnelt, M. G. 2004 Joint Formation of Supermassive Black Holes and Galaxies. In Coevo- lution of Black Holes and Galaxies (ed. L. C. Ho), p. 406. Cambridge University Press. Hogg, D. W. 2001 A meta-analysis of cosmic star-formation history submitted to PASP (astro-ph/0105280) Hopkins, A. M. 2004 On the Evolution of Star Forming Galaxies ApJ in press. (astro-ph/0407170) Im, M., et al. 2002 The DEEP Groth Strip Survey X. Number Density and Luminosity Function of Field E/S0 Galaxies at z < 1 ApJ 571, 136 -- 171. Kauffmann, G., et al. 2003 Stellar masses and star formation histories for 105 galaxies from the Sloan Digital Sky Survey MNRAS 341, 33 -- 53. Kelson, D. D., Illingworth, G. D., Franx, M., & van Dokkum, P. G. 2001 The Evolution of Balmer Absorption-Line Strengths in E/S0 Galaxies from z = 0 to z = 0.83 ApJ 522, L17 -- 21. Kennicutt, Jr., R. C. 1983 The rate of star formation in normal disk galaxies ApJ 272, 54 -- 67. Kennicutt, Jr., R. C. 1998 Star Formation in Galaxies Along the Hubble Sequence ARA&A 36, 189 -- 232. Kodama, T., et al. 2004 Down-sizing in galaxy formation at z ∼ 1 in the Subaru/XMM- Newton Deep Survey (SXDS) MNRAS 350, 1005 -- 1014. Kong, X., Charlot, S., Brinchmann, J., Fall, S. M. 2004 Star formation history and dust content of galaxies drawn from ultraviolet surveys MNRAS 349, 769 -- 778. Kroupa, P. 2001 On the variation of the initial mass function MNRAS 322, 231 -- 246. Larson, R. B. 1974 Dynamical models for the formation and evolution of spherical galaxies MNRAS 166, 585 -- 616. Le F`evre, O., et al. 2000 Hubble Space Telescope imaging of the CFRS and LDSS redshift surveys -- IV. Influence of mergers in the evolution of faint field galaxies from z ∼ 1 MNRAS 311, 565 -- 575. Lilly, S. J., Tresse, L., Hammer, F., Crampton, D., & Le F`evre, O., 1995 The Canada- France Redshift Survey: VI. Evolution of the Galaxy Luminosity Function to z ∼ 1 ApJ 455, 108 -- 124. Lilly, S. J., Le F`evre, O., Hammer, F., & Crampton, D. 1996 The Canada-France Redshift Survey: The Luminosity Density and Star Formation History of the Universe to z ∼ 1 ApJ 460, L1-4 Lisenfeld, U., Volk, H. J., & Xu, C. 1996 The FIR/radio correlation in starburst galaxies: constraints on starburst models A&A 314, 745 -- 753. Lotz, J. M., Primack, J., & Madau, P. 2004 A New Nonparametric Approach to Galaxy Morphological Classification AJ 128, 163 -- 182. Madau, P., Ferguson, H. C., Dickinson, M. E., Giavalisco, M., Steidel, C. C., & Fruchter, A. 1996 High-redshift galaxies in the Hubble Deep Field: colour selection and star formation history to z ∼ 4 MNRAS 283, 1388 -- 1404. Meza, A., Navarro, J. F., Steinmetz, M., & Eke, V. R. 2003 Simulations of Galaxy For- mation in a ΛCDM Universe III. The Dissipative Formation of an Elliptical Galaxy ApJ 590, 619 -- 635. Moustakas, L. A., et al. 2004 Morphologies and Spectral Energy Distributions of Extremely Red Galaxies in the GOODS-South Field ApJ 600, L131 -- 134. Naab, T., & Burkert, A. 2003 Statistical Properties of Collisionless Equal- and Unequal-Mass Merger Remnants of Disk Galaxies ApJ 597, 893 -- 906. Niklas, S., & Beck, R. 1997 A new approach to the radio -- far infrared correlation for non- calorimeter galaxies A&A 320, 54 -- 64. 22 E. F. Bell: Galaxy Assembly Papovich, C., Dickinson, M. E., & Ferguson, H. C. 2001 The Stellar Populations and Evolution of Lyman Break Galaxies ApJ 559, 620 -- 653. Patton, D. R., et al. 2000 New Techniques for Relating Dynamically Close Galaxy Pairs to Merger and Accretion Rates: Application to the Second Southern Sky Redshift Survey ApJ 536, 153 -- 172. Patton, D. R., et al. 2002 Dynamically Close Galaxy Pairs and Merger Rate Evolution in the CNOC2 Redshift Survey ApJ 565, 208 -- 222. Peebles, P. J. E. 1980 The large-scale structure of the universe. Princeton, N. J.: Princeton University Press. Riess, A. G., et al. 2004 Type Ia Supernova Discoveries at z > 1 from the Hubble Space Telescope: Evidence for Past Deceleration and Constraints on Dark Energy Evolution ApJ 607, 665 -- 687. Rix, H.-W., et al. 2004 GEMS: Galaxy Evolution from Morphology and SEDs ApJS 152, 163 -- 173. Rudnick, G., et al. 2003 The Rest-Frame Optical Luminosity Density, Color, and Stellar Mass Density of the Universe from z = 0 to z = 3 ApJ 599, 847 -- 864. Sandage, A., & Visvanathan, N. 1978 The color-absolute magnitude relation for E and S0 galaxies. II -- New colors, magnitudes, and types for 405 galaxies ApJ 223, 707 -- 729. Schweizer, F., & Seitzer, P. 1992 Correlations between UBV colors and fine structure in E and S0 galaxies -- A first attempt at dating ancient merger events AJ 104, 1039 -- 1067. Seljak, U., et al. 2004 Cosmological parameter analysis including SDSS Ly-alpha forest and galaxy bias: constraints on the primordial spectrum of fluctuations, neutrino mass, and dark energy submitted to PRD (astro-ph/0407372). S´ersic, J. L. 1968 Atlas de Galaxias Australes. Cordoba: Observatorio Astronomico. Skrutskie, M. F., et al. 1997 The Two Micron All Sky Survey (2MASS): Overview and Status. in The Impact of Large Scale Near-IR Sky Surveys (eds. F. Garzon et al.) p. 25. Dordrecht: Kluwer Academic Publishing Company Somerville, R. S., Lee, K., Ferguson, H. C., Gardner, J. P., Moustakas, L. A., & Giavalisco, M. 2004 Cosmic Variance in the Great Observatories Origins Deep Survey ApJ 600, L171 -- 174. Spergel, D. N., et al. 2003 First-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Determination of Cosmological Parameters ApJS 148, 175 -- 194. Stanway, E. R., Bunker, A. J., McMahon, R. G., Ellis, R. S., Treu, T., & McCarthy, P. J. 2004 Hubble Space Telescope Imaging and Keck Spectroscopy of z ∼ 6 i-Band Dropout Galaxies in the Advanced Camera for Surveys GOODS Fields ApJ 607, 704 -- 720. Steidel, C. C., Adelberger, K. L., Giavalisco, M., Dickinson, M., & Pettini, M. 1999 ∼ 4 and the Evolution of the Ultraviolet Luminosity Density Lyman-Break Galaxies at z > at High Redshift ApJ 519, 1 -- 17. Strateva, I., et al. 2001 Color Separation of Galaxy Types in the Sloan Digital Sky Survey Imaging Data AJ 122, 1861 -- 1874. Toomre, A., & Toomre, J. 1972 Galactic Bridges and Tails ApJ 178, 623 -- 666. Tully, R. B., Pierce, M. J., Huang, J. S., Saunders, W., Verheijen, M. A. W., & Witchalls, P. L. 1998 Global Extinction in Spiral Galaxies AJ 115, 2264 -- 2272. van Kampen, E., Jimenez, R., & Peacock, J. A. 1999 Overmerging and mass-to-light ratios in phenomenological galaxy formation models MNRAS 310, 43 -- 56. White, S. D. M., & Frenk, C. S. 1991 Galaxy formation through hierarchical clustering ApJ 379, 52 -- 79. White, S. D. M., & Rees, M. J. 1978 Core condensation in heavy halos -- A two-stage theory for galaxy formation and clustering MNRAS 183, 341 -- 358. Witt, A. N., & Gordon, K. D. 2000 Multiple Scattering in Clumpy Media II. Galactic Environments ApJ 528, 799 -- 816. Wolf, C., Meisenheimer, K., Rix, H.-W., Borch, A., Dye, S., & Kleinheinrich, M. 2003 The COMBO-17 survey: Evolution of the galaxy luminosity function from 25 000 galaxies with 0.2 < z < 1.2 A&A 401, 73 -- 98. Worthey, G. 1994 Comprehensive stellar population models and the disentanglement of age and metallicity effects ApJS 95, 107 -- 149. Wyuts, S., van Dokkum, P. G., Kelson, D. D., Franx, M., Illingworth, G. D. 2004 E. F. Bell: Galaxy Assembly 23 The Detailed Fundamental Plane of Two High-Redshift Clusters: MS 2053-04 at z = 0.58 and MS 1054-03 at z = 0.83 ApJ 605, 677 -- 688. Yan, L., & Thompson, D. 2003 Hubble Space Telescope WFPC2 Morphologies of K-selected Extremely Red Galaxies ApJ 586, 765 -- 779. Yee, H. K. C., & Ellingson, E. 1995 Statistics of close galaxy pairs from a faint-galaxy redshift survey ApJ 445, 37 -- 45. York, D. G., et al. 2000 The Sloan Digital Sky Survey: Technical Summary AJ 120, 1579 -- 1587. Zepf, S. E., & Koo, D. C. 1989 Close pairs of galaxies in deep sky surveys ApJ 337, 34 -- 44.
astro-ph/0009208
3
0009
2001-08-10T02:14:03
Modeling Mid-Ultraviolet Spectra. I. Temperatures of Metal-Poor Stars
[ "astro-ph" ]
Determining the properties of old stellar systems using evolutionary population synthesis requires a library of reliable model stellar fluxes. Empirical libraries are limited to spectra of stars in the solar neighborhood, with nearly solar abundances and abundance ratios. We report here a first step towards providing a flux library that includes nonsolar abundances, based on calculations from first principles that are calibrated empirically. We have started with main-sequence stars, whose light dominates the mid-ultraviolet spectrum of an old stellar system. We have calculated mid-ultraviolet spectra for the Sun and nine nearby, near-main-sequence stars spanning metallicities from less than 1/100 solar to greater than solar, encompassing a range of light-element abundance enhancements. We first determined temperatures of eight of the stars by analyzing optical echelle spectra together with the mid-ultraviolet. Both could be matched at the same time only when models with no convective overshoot were adopted, and only when an approximate chromosphere was incorporated near the surface of relatively metal-rich models. Extensive modifications to mid-UV line parameters were also required, notably the manual assignment of approximate identifications for mid-UV lines missing from laboratory linelists. Without recourse to additional missing opacity, these measures suffice to reproduce in detail almost the entire mid-UV spectrum of solar-temperature stars up to one-tenth solar metallicity, and the region from 2900A to 3100A throughout the entire metallicity range. Ramifications for abundance determinations in individual metal-poor stars and for age-metallicity determinations of old stellar systems are briefly discussed, with emphasis on the predictive power of the calculations.
astro-ph
astro-ph
To appear in The Astrophysical Journal, September 20, 2001 Modeling Mid-Ultraviolet Spectra. I. Temperatures of Metal-Poor Stars1,2 Ruth C. Peterson3, Ben Dorman4, and Robert T. Rood5 ABSTRACT Determining the properties of remote globular clusters and elliptical galaxies using evolutionary population synthesis requires a library of reliable model stellar fluxes. Em- pirical libraries are limited to spectra of stars in the solar neighborhood, with nearly solar abundances and abundance ratios. We report here a first step towards providing a flux library that includes nonsolar abundances, based on calculations from first prin- ciples that are calibrated empirically. Because the mid-ultraviolet spectrum of an old stellar system is dominated by the contribution from its main-sequence turnoff stars, we have started by modeling these. We have calculated mid-ultraviolet spectra for the Sun and nine nearby, near-main-sequence stars spanning metallicities from less than 1/100 solar to greater than solar, encompassing a range of light-element abundance enhancements. We first determined temperatures of eight of the stars by analyzing optical echelle spectra together with the mid-ultraviolet. Both could be matched at the same time only when models with no convective overshoot were adopted, and only when an approxi- mate chromosphere was incorporated near the surface of relatively metal-rich models. Extensive modifications to mid-UV line parameters were also required, notably the man- ual assignment of approximate identifications for mid-UV lines missing from laboratory linelists. Without recourse to additional missing opacity, these measures suffice to re- produce in detail almost the entire mid-UV spectrum of solar-temperature stars up to one-tenth solar metallicity, and the region from 2900 A to 3100 A throughout the entire metallicity range. Ramifications for abundance determinations in individual metal-poor stars and for age-metallicity determinations of old stellar systems are briefly discussed, with emphasis on the predictive power of the calculations. 1Based on observations obtained with the Hubble Space Telescope of Space Telescope Science Institute, under contract with the National Aeronautics and Space Administration (NASA). 2Based on observations obtained with the Shane Telescope at Mt. Hamilton, UCO/Lick Observatory. 3UCO/Lick Observatory, Dept. of Astronomy, University of California, Santa Cruz, CA 95064, and Astrophysical Advances, Palo Alto, CA 94301; [email protected]. 4Emergent-IT Inc. and Goddard Space Flight Center, Greenbelt, MD 20771; [email protected]. 5Dept. of Astronomy, University of Virginia, P.O. Box 3818, Charlottesville, VA 22903-0818; [email protected]. -- 2 -- Subject headings: stars: abundances -- stars: fundamental parameters -- stars: Popu- lation II -- galaxies: abundances -- ultraviolet: galaxies -- ultraviolet: stars -- 3 -- 1. Introduction To characterize the metallicity and age of an old stellar system, iron abundance [Fe/H] (the logarithm of the iron-to-hydrogen abundance ratio with respect to its solar value) and the stellar effective temperature Teff must be determined together for its main-sequence turnoff (MSTO) stars. These are dwarfs and subgiants of types F and early G with 5750 ≤ Teff ≤ 7000 K. Their spectra are dominated by Fe i lines, whose strength increases as Teff declines. For both metallicity and age, then, a reliable derivation of Teff is necessary. This can be problematical even for an individual star. The photometric colors often used to derive Teff are sensitive to reddening by the interstellar medium and to the modeling of convection (Castelli, Gratton, & Kurucz 1997), as is the more recently developed infrared flux method (Alonso, Arribas, & Martinez-Roger 1996). Methods using Balmer-line profiles are reddening-independent but still very model-dependent (Magain 1984; Fuhrmann, Axer, & Gehren 1994). Methods relying on the excitation equilibrium of Fe i are reddening-independent and model-independent if lines are weak, but are susceptible to systematic errors in laboratory measurements of gf-values (Blackwell et al. 1982) and to possible departures from local thermodynamic equilibrium (LTE; Th´evenin & Idiart 1999). Currently, Teff values derived for individual MSTO stars show uncomfortably large spreads: values of Carney et al. (1994) and Fulbright (2000) average 100 K -- 200 K lower than those found by King (1993) and Gratton, Carretta, & Castelli (1996), with those of Alonso, Arribas, & Martinez- Roger (1996) and Blackwell & Lynas-Gray (1998) in between. Not only is overall metallicity affected, but also the ratios of the abundance with respect to iron of certain elements. Among these are the [Li/Fe] and [O/Fe] ratios so critical to uncovering the nucleosynthesis history of the early universe (Spite et al. 1996) and deducing the age of globular clusters from the MSTO luminosity (VandenBerg & Bell 2000). When Teff alone is raised by 100 K, modeled values of [Li/Fe] change by +0.09 dex (Spite et al. 1996); [O/Fe] changes by −0.09 dex, +0.03 dex, or +0.16 dex depending on whether the high-excitation O i 7775 A triplet, the [O i] 6300 A ground-state line, or OH molecular lines near 3100 A are used (Kraft 2000). Not surprisingly, the oxygen abundance in extremely metal-poor stars remains uncertain, as summarized during a workshop devoted to the problem (Kraft 2000; Balachandran 2000; Peterson 2000; Lambert 2000). King (1993) and Kraft (2000), among others, have suggested that a higher Teff scale for metal-poor turnoff stars might eliminate the systematic differences in [O/Fe] seen from its different diagnostics. For all these reasons, it seems important to reduce uncertainties in the determination of Teff in MSTO stars. Fitting the slope of their UV flux distribution might help, since the fluxes of F -- G stars peak in the optical so that UV fluxes vary rapidly with Teff . Although fitting a UV slope is also reddening-dependent, reddening should be very low for the brightest metal-poor MSTO stars, which are typically within 50 pc (e.g., Reid 1998). Reliable calculations of MSTO mid-UV spectra would be especially valuable in constraining the age and metallicity of a spatially unresolved old stellar system such as an elliptical galaxy or -- 4 -- extragalactic globular cluster. For the mid-UV flux of an old stellar population is dominated by turnoff stars (Dorman, Rood, & O'Connell 1993), but the light redward of 4500 A is dominated by cool giants in a 17 Gyr population of solar metallicity (Worthey 1994). To date, however, mid-UV studies of galaxies and globulars (e.g., Spinrad et al. 1997; Ponder et al. 1998) have largely been empirical, comparing extragalactic spectra against observed spectra of individual F and G stars (Fanelli et al. 1990). Thus the metallicity of the extragalactic systems has generally been assumed to be solar, although age is very sensitive to this assumption (Heap et al. 1998). Furthermore, nonsolar abundance ratios have not been included, despite the fact that abundance enhancements of light elements such as magnesium are generally seen in metal-rich extragalactic systems (Peterson 1976; O'Connell 1976; Worthey, Faber, & Gonzalez 1992; Henry & Worthey 1999). Modeling the mid-UV region is difficult, though, because of the dramatic increase in line absorption at short wavelengths. Line blending in the solar spectrum suppresses the optical con- tinuum blueward of ∼ 4500 A (Kurucz et al. 1984), and becomes even more severe in the near- and mid-UV. Especially in the UV, many lines are "missing" -- although they appear as absorption features in spectra of solar-type stars, they lack laboratory identification, and are thus not matched by spectral synthesis calculations (Kurucz & Avrett 1981). As a result, the normalization of the solar UV continuum is controversial (Balachandran & Bell 1998; Boesgaard et al. 1999; Israelian et al. 2001). With few exceptions (e.g., Heap et al. 1998), calculations of the mid-UV region are based on Kurucz model flux distributions. Allende Prieto & Lambert (2000) compared these to International Ultraviolet Explorer (IUE) mid-UV spectra of nearby metal-poor halo stars, illustrating that Teff is generally recovered when parallaxes are known and reddening negligible. Lotz, Ferguson, & Bohlin (2000) have used Kurucz fluxes as generated by Lejeune, Cuisinier, & Buser (1997) to examine the effect of abundance on the mid-UV light of old stellar populations, but found that the UV fluxes of cool stars were poorly reproduced. This is not surprising given the known limitations of the opacity distribution functions on which the Kurucz flux distributions are based. They statistically treat lines in a list that includes "predicted" lines whose wavelengths are uncertain by ∼10 A (because one or both energy levels have not been measured in the laboratory), limiting the resolution of the flux distributions to 10 A. Moreover, they assume a solar iron abundance ∼0.15 dex higher than the value currently accepted (Bi´emont et al. 1991; Holweger, Kock, & Bard 1995), and a solar ratio of light-element abundances [α/Fe] = 0. In contrast, metal-poor stars of the Galactic halo typically show [α/Fe] = 0.3 (Wheeler, Sneden, & Truran 1989), a factor-of-two enhancement in Mg, Si, Ca, and Ti. This affects the fit in the mid-UV, since it includes the very strong resonance doublet of Mg ii at 2800 A and a bound-free opacity edge of Mg i at 2512 A, plus many weaker Mg, Si, and Ti lines. -- 5 -- 2. Overview of This Work In this work we calculate the mid-UV spectrum line-by-line at high resolution, independent of the Kurucz flux distributions. Our line list does not include predicted lines, but rather is based on an updated laboratory line list augmented by manual addition of missing lines where necessary. To ensure the reliability of these calculations, we match as best we can the individual line blends observed in the mid-UV echelle spectra of real near-turnoff stars, spanning a wide metallicity range and including both solar and enhanced light-element ratios. Our purpose is to reproduce mid-UV spectra of turnoff stars accurately enough to determine Teff and to predict the composite mid- UV spectra of old stellar systems over their entire range of overall metallicity and light-element abundance ratio. In this first step, we describe this calibration procedure. Because the calibrations depend significantly on the stellar temperatures adopted, we have carried out new Teff determinations in seven metal-poor turnoff stars and one mildly metal-poor hotter star. They are based on simulta- neously matching both mid-UV spectra (2280 -- 3120 A ) and the Hα profile and the strengths of high-excitation atomic lines in optical spectra. We thus do not rely on any photometry or parallax information in the Teff determination. Iron abundances and [α/Fe] ratios were adopted as derived from both weak and strong lines in the optical spectra. To assist in filling in the mid-UV line list, we have included mid-UV analyses of two stars with mid-UV spectra of very high resolution, the Sun and HD 128620 (= α Cen A ), a well-studied nearby southern star with a greater-than-solar metallicity (Neuforge-Verheecke & Magain 1997), for which Teff was taken from the literature. In fitting each mid-UV spectrum, we adopted a single continuum normalization constant, the scale factor by which the observed mid-UV spectral flux was multiplied to match the flux predicted by theoretical calculation. This constant proved to agree to an average of 6% ± 5% with that expected from the ratio of stellar distance (found from the parallax) to stellar radius (found from the model log g assuming a reasonable stellar mass). Model U − B , B − V , and b − y colors all agreed extremely well with those observed, confirming Teff errors of .50 K. Our mid-UV and optical echelle spectra are described in §3. As discussed in §4, Teff was initially established from the optical spectra, and compared to the mid-UV fit. For the most metal-poor stars, this fit did not require the addition of missing lines, but did require the use of Kurucz models in which convective overshoot was turned off, in contrast to the Kurucz models in general use. To reproduce mid-UV line strengths and to match the pseudocontinuum of the mid-UV spectra of more metal-rich stars, mid-UV atomic line parameters were changed and missing lines were added. Transitions especially sensitive to the surface temperature were still poorly reproduced; this mismatch was reduced by adopting a mild enhancement of temperatures in shallow photospheric layers of the models of one-tenth solar metallicity and higher. The degree to which the resulting calculations match both mid-UV and optical spectra of turnoff stars is shown and discussed in §5. The match is very good for 2900 -- 3120 A at all metallicities, and generally good throughout the mid-UV for metallicities at and below one-third -- 6 -- solar. It could be improved at solar metallicities by obtaining additional high-resolution mid-UV spectra of a few well-chosen standards. Our Teff results are presented in §6, where comparisons are made of model fluxes and colors with observations. We conclude in §7 with a summary and a discussion of the prospects for constraining age and metallicity of quiescent stellar systems by comparing such calculations to spectra of their integrated light. 3. Observations We begin by listing the properties of the stars considered in Table 1. The first five columns following the HD number give the observed V magnitude, colors, and the parallax from SIMBAD. The next five columns, discussed in §6, show comparisons of observed properties of each star with those calculated from the models. The parameters of the model used to calculate both mid-UV and optical spectra of each star are given in the next four columns. In all cases but HD 128620 = α Cen A and the Sun, all the model parameters were redetermined here as described below. The final columns show Teff values determined previously in several works. Of the eight stars whose temperatures have been redetermined here from optical and UV spectra, seven are metal-poor main-sequence turnoff (MSTO) stars. The eighth, HD 128167, is a mildly-metal poor star lying above the halo main-sequence turnoff, either younger or a blue straggler. It was included to give temperature leverage to the mid-UV comparisons, as discussed below. Likewise the Sun and α Cen A were included to represent the high-metallicity end. The solar parameters were considered so well determined that no re-evaluation was necessary. The star α Cen A was not reanalyzed because no optical spectrum was available: it is inaccessible from the northern hemisphere. Its Teff and abundance have been determined to moderate accuracy from extensive abundance analyses, and its gravity as log g = 4.33 ± 0.02 as found from its visual and radial-velocity motion about its companion (Chmielewski et al. 1992; Neuforge-Verheecke & Magain 1997; Pourbaix et al. 1999). The iron abundance listed in Table 1 was adopted to best match its mid-UV spectrum. Optical spectra were obtained for all stars in Table 1 but the Sun and α Cen A . The Lick Observatory Shane 3m echelle spectrograph (Vogt 1987) was used with a TEK 2048 × 2048 CCD, at a FWHM resolution of 38,000. The spectra of HD 84937, HD 94028, and HD 106516 were kindly provided by Fulbright (2000). Exposure times on the 3m were typically 10min for all stars except HD 128167 (40 s). Complete spectral coverage over 3900 A -- 7800 A and S/N > 100/pixel were obtained in all cases but one, although fringing compromises the accuracy of the data at the long- wavelength end of the range. For the one exception, HD 114762, the blue spectrum was obtained in two one-hour exposures by Tony Misch at Lick with the coude auxiliary telescope, using the same echelle spectrograph configuration. For the Sun, the flux spectrum of Kurucz et al. (1984) was downloaded from the Kurucz web site at http://cfaku5.harvard.edu; the full-resolution version -- 7 -- was matched at 150,000 resolution. The mid-UV spectral observations of the Sun modeled here are the rocket spectra of Kurucz & Avrett (1981), recorded on film. Both passes of the center of the solar disk were used. Gaussian broadenings of 1.5 km s−1 for macroturbulence and 80,000 FWHM for other broadening sources provided a good match. A constant normalization scale factor of 1.10 was adopted, in keeping with the stated 10% uncertainty of its flux calibration. The stellar UV spectra were obtained with the Space Telescope Imaging Spectrograph (STIS) of Hubble Space Telescope (HST) from mid-1998 to mid-1999. Salient features are noted here; full details may be found from the HST archive at http://archive.stsci.edu. To obtain the mid-UV spectra of most stars, the near-UV MAMA detector was used with the STIS E230M grating over its default range of 2280 A -- 3120 A at a FWHM resolution ∼25,000. HD 106516 (= HR 4657) was observed by Heap et al. (1998) in program #7433 with the 0.2 × 0.06 aperture. Mid-UV spectra for HD 19445, HD 84937, HD 94028, HD 114762, HD 184499, and HD 201891 were acquired in snapshot program #7402 by Peterson & Schrijver (2000), with two exposures obtained for HD 19445, HD 84937, and HD 114762. The 0.2 × 0.2 aperture was used and exposure times were 6 -- 10 min. Consequently, S/N is lower, but flux levels are less dependent on pointing. Fluxes good to 1% for these six stars are indicated by previous observations of Mg ii line profiles in each star, made using the HST Goddard High-Resolution Spectrograph (GHRS) and the 2 × 2 aperture in program #5869 (Peterson & Schrijver 1997). Mid-UV spectra of the stars HD 128167 (=HR 5447), from #7433, and HD 128620 (= HR 5459 =α Cen A), from #7263 (Linsky et al. 2000), were recorded in multiple exposures with the E230H grating. Spectral coverage is 2380 A -- 3160 A for HD 128167 and <2280 A -- 3160 A for HD 128620. A FWHM resolution of ∼60,000 was achieved with the 0.2 × 0.06 aperture for HD 128167, and the 0.2 × 0.05ND aperture for HD 128620. For neither star do fluxes always reproduce exactly in orders obtained in more than one exposure; differences as large as 10% are seen. Similar uncertainties are anticipated for HD 106516, where the 0.2 × 0.06 aperture was also used. Rotation is detected in at least two stars and possibly in two more, from the extra broadening required to match line profiles. For HD 128167, HD 106516, HD 84937, and HD 114762, line breadths were matched in the spectral synthesis by adopting rotational velocities v sin i = 9, 7, 5, and 4 km s−1 respectively. Thus the resolution of the mid-UV spectrum of HD 128167 is effectively degraded to that of the stars observed at lower resolution, affecting the discernment of mid-UV lines. This star nonetheless has one of the sharpest-lined spectra among those obtained in program #7433. 4. Spectrum Analysis Spectrum analysis was based on visually matching each observed spectrum to ab initio radiative- transfer calculations assuming static models in LTE. The program SYNTHE (Kurucz & Avrett -- 8 -- 1981), the Kurucz (1995) and Castelli et al. (1997) grids of ATLAS9 models, and the Kurucz line lists for atomic species and molecular hydrides were downloaded from the Kurucz web site, where further details on each of these can be found. As described below, all models, including that of the Sun, were ultimately drawn from the Castelli et al. (1997) grid alone. The more metal-rich ones were modified by introducing enhanced temperatures in the shallow layers. The SYNTHE program accepts as input both line lists and a model atmosphere with tempera- ture and other quantities tabulated for each depth, and calculates an array of spectra at individual emergent angles. These may then be used individually (as for spectra of the center of the solar disk) or convolved with a zero or positive rotational velocity to generate an emergent flux spectrum. In either case, the resulting spectrum can be broadened by Gaussians (or other profile shapes) to incorporate macroturbulence and instrumental broadening. SYNTHE was modified to run under UNIX on a Sun Ultra-30 at the University of Virginia. Mid-UV spectra from 2280 A to 3160 A were calculated at a resolution of 330,000 (500,000 for the Sun and HD 128620), and optical spectra at a resolution of 500,000. A Gaussian macroturbulent broadening of 1.5 km s−1 was assumed for all stars as well as for the Sun. Spectra for some stars were rotationally broadened as listed in the last paragraph of §3, as was the solar flux spectrum, for which a mean rotational velocity of 2 km s−1 was adopted. The spectra were then broadened by Gaussians corresponding to the values of instrumental resolution listed individually in §3. The line lists and the source code of the opacity routines available at the Kurucz web site give specifics on the wide variety of opacity sources treated by SYNTHE. For example, the significant absorption by the bound-free absorption edge of Mg i near 2512 A is treated both as line opacity, with the line list including the bound-bound transitions to N ∼ 60, and as continuous opacity blueward of the bound-free edge at νo = 39759.8 cm−2, by the subroutine MG1OP. Near νo, this routine adopts a cross section of 40 Mbarn with a (ν/νo)−14 frequency dependence, based on the laboratory measurement of 46 ± 12 Mbarn by Lombardi, Smith, & Parkinson (1981) and the convergence of high-level lines. Where this formalism falls below the theoretical cross section, the program adopts the latter, with a cross section of 20 Mbarn and a (ν/νo)−2.7 frequency dependence (e.g., Butler et al. 1993). For calculations redward of 5000 A, the line list of Peterson, Dalle Ore, & Kurucz (1993) based on the solar spectrum was adopted, with modifications due to the use here of a flux, rather than central intensity, spectrum for the Sun and to the incorporation of an approximate chromosphere in the solar model. This led to the deduction of larger gf-values for the strongest lines except where laboratory gf-values were adopted, as for low-excitation Fe i lines with furnace measurements (Blackwell et al. 1982). Blueward of 5000 A , the Kurucz 25 May 1998 atomic line lists gf 0300.100, gf 0400.100, and gf 00500.100 were used. These differ significantly from those used in the Kurucz flux distribution calculations in including only atomic lines identified in the laboratory (with laboratory gf-values where available) -- like those of Kurucz & Bell (1995) but with the addition of many lines of Fe i, Fe ii, -- 9 -- and other species whose upper energy levels were newly determined in near-UV laboratory spectra (e.g., Nave et al. 1994). Because the strengths or profiles of many lines were mismatched when using line parameters directly from this list, their gf-values and damping constants were changed where necessary, as described shortly. Since many lines appear in the STIS mid-UV spectra that were not modeled by lines in the list, such missing lines were added blueward of 3120 A. Redward of this, no missing lines were added, but gf-values were changed to match stronger lines in the Sun in the regions depicted in Fig. 4 below. The Teff , log g, [Fe/H], χt, and [α/Fe] values for each reanalyzed star were initially established from Hα and weak lines of neutral species in the optical spectra, reassessed from optical lines of once-ionized species, then confirmed or revised from the UV. For the two most metal-poor stars, mid-UV spectra provided an immediate check on Teff . For the others, the following steps were iterated until all diagnostics converged. First the Hα profile and an assumed log g were used to estimate Teff . [Fe/H] was adjusted until weak Fe i lines in the 5100 A -- 5400 A region were well matched; their strengths are largely gravity- independent. The strength of weak Ti i, Mg i, and Si i lines determined [α/Fe] . Microturbulent velocity χt was set by matching weak and strong iron and titanium lines. Gravity log g was then checked by demanding matches for Mg i, Mg ii, and the wings of the Mgb lines, for Fe i and Fe ii, for Ti i and Ti ii, and for Si i and Si ii. This also constrains Teff , for the Mg ii and Si ii doublets arise from highly excited lower levels (>8 eV), while the optical Ti ii and Fe ii lines originate from much lower levels (<3 eV). Roughly two dozen Fe i lines, a dozen Ti i lines, and a dozen total Ti ii and Fe ii lines were detected for the most metal-poor stars, with double this number typical of the more metal-rich stars. The Mg ii and Si ii doublets at 4481 A and at 6347.1 and 6371.3 A were seen in all cases, but Si ii only marginally so in HD 84937 and HD 19445. Once the diagnostics began to converge, iterative adjustments were made to the Kurucz mid- UV atomic-line parameters. First, gf-values were modified for atomic lines identified in the labora- tory. The relative contribution of identified lines to a blend was ascertained from the behavior of the blend in the hot star HD 128167 versus the cooler ones, and in the higher-gravity stars such as HD 19445 versus the lower-gravity ones. Damping constants were changed in some cases to match line profiles. Damping constants rather than gf-values were changed for the strong resonance lines well-observed in the interstellar medium for which gf-values are tabulated by Morton (1991). Next, for lines lacking identification, the laboratory line list was searched for transitions close enough in wavelength to perhaps be responsible. The gf-value of a nearby identified line was increased by up to 4.5 in the log, and the calculation repeated. This line was dropped if still not strong enough; otherwise its gf-value was adjusted to match. The high resolution of the α Cen A and solar spectra showed that most such identifications were very well matched in wavelength, and thus likely to be correct, when based on increases of ≤ 2.5 in the log. For the two most metal-poor stars, these steps were sufficient to enable the continuum to be defined unambiguously throughout most of the mid-UV region. Its definition in more metal- -- 10 -- rich stars was hampered by line crowding, yielding a pseudocontinuum that the modeling did not properly reproduce. Much of this pseudocontinuum mismatch appeared to be due to individual transitions missing from the line list. This was indicated by the very common occurrence throughout the mid-UV of unmodeled absorption features found at the same wavelengths in all stars, features whose strength steadily increased in going from weak-lined to strong-lined spectra. Most of these transitions are weak, consistent with the fact that the intrinsically strongest transitions of each species are the ones most likely to have been detected in the laboratory. Except just redward of the Mg i edge at 2512 A, most missing lines are probably due to iron. Its abundance is high and the spectra are rich in lines of Fe i and Fe ii, which are well-populated species in solar-temperature stars. Of the elements with similar first ionization potentials, only magnesium and silicon have comparable abundances; the mid-UV spectra of their low-ionization stages are generally much less complex, as are those of the more abundant CNO elements. While iron-peak elements such as Cr, Mn, and Ni have many transitions throughout the mid-UV, iron dominates because it is twenty or more times as abundant. In most halo stars above one-tenth solar metallicity, the relative abundances of these elements with respect to iron are within 0.2 dex of those of the Sun (although larger differences may occur, especially at very low metallicity: McWilliam et al. 1995; Nissen et al. 2000; Prochaska & McWilliam 2000). The mismatch that will result from assigning missing lines of these elements to iron should be mild. In the region immediately redward of the bound-free Mg i opacity edge at 2512 A, the proba- bility is significant that missing lines may be due to magnesium rather than iron. Redward of the edge there are a very large number of transitions to known highly excited levels of the Mg i atom, and presumably a very large number of similar transitions to closely-spaced highly excited levels If many weak transitions redward of 2512 A are that are as yet unidentified in the laboratory. due to magnesium but have been assigned to iron instead, our calculations will underestimate the absorption in this region in metal-rich alpha-enhanced populations. We see below that this region is poorly modeled in any case once metallicity exceeds one-tenth solar. The modeling then proceeded iteratively by assigning tentative identifications and transition probabilities to each missing feature, then recalculating until an acceptable match was achieved or deemed unattainable with existing data (as for the regions just redward of 2550 A and 2640 A in the solar spectrum). Features were attributed to either low-excitation Fe i, moderate-excitation Fe i, or Fe ii, whose line strengths depend differently on stellar temperature and gravity (pressure) because of the Boltzmann and Saha equations. With a Teff ∼1000 K higher than those of other moderately weak-lined stars, HD 128167 proved exceptionally useful in making these assignments. The higher resolution of the spectra of the strong-lined stars HD 128620 = α Cen A and the Sun was essential in discerning the wavelengths of the many features emerging above one-third solar metallicity. A brief examination was made of the gf-value changes for low-excitation Fe i lines, namely those identified in the laboratory with (air) wavelengths falling between 2600 A and 3000 A , and -- 11 -- with a lower level of 6000 -- 10000 cm−1 and an upper level of 40000 -- 50000 cm−1. Sixty-five such Fe i lines have gf- values from the laboratory measurements of O'Brian et al. (1991); 52 were not changed, while 13 were changed by from −0.6 to +0.4 dex. Nine lines have gf-values from the laboratory measurements of Fuhr, Martin, & Wiese (1988); five were unchanged, and four changed by from −0.7 to −0.3 dex. Thus the mid-UV gf-values found in this work generally concur with the mid-UV laboratory measurements. In contrast, fifteen lines have theoretical gf-values from Kurucz (1994); four were not changed, but 11 were changed by from −2.2 to +0.6 dex, all but one decreased. The theoretical gf-values of Kurucz (1994) are seen to be subject occasionally to very large errors, usually overestimates. This can be attributed to the finite size of the Cray computer used in those calculations, which dictated that the number of levels included be truncated, and so upper levels were preferentially omitted from the model atom. Because the sum of the gf-values of transitions arising from a given lower state is normalized to unity, gf-values for the transitions included in the calculation will be overestimated because of the exclusion of others arising from the same level. Thus the calculations will overestimate the gf-value by increasing amounts as the fraction decreases of the actual possible upper states that are included in the atomic model. This fraction decreases to the blue, because higher upper levels are involved for a given lower level. Indeed, the severity of overestimates appears to increase to the blue: all the above changes made to the Kurucz (1994) gf-values were −1.5 or lower below 2650 A, but redward of this only one change this negative was required. The overestimates among the Kurucz (1994) Fe i gf-value calculations are thus seen to be due to incompleteness and not to more fundamental difficulties. Undoubtedly many of the weaker mid-UV transitions are missing entirely from the predicted list. For the hydrides, separate treatment was required, as the lines are naturally associated into bands, and the Kurucz hydride calculations are now several decades old. For OH, the theoretical calculations of Gillis et al. (2001) (communicated in August 2000 by M. Bessell) were used wherever possible. For CH, the LIFBASE data (e.g., Luque & Crosley 1999) were used (communicated in August 2000 by M. Bessell with lines cross-identified). Since these cover only the strongest bands, calculations must still rely on the Kurucz values for the others, notably the 4-3 band of OH with many weak but discernible lines in the Sun. A comparison of the Kurucz OH transition probabilities against those of Gillis et al. (2001) showed strong trends, with discrepancies increasing at high rotation number J. To place them on approximately the same scale, the Kurucz transition probabilities for the OH and CH lines were decreased by an amount δ = −0.25 − 0.05 ∗ Jl dex. Once missing atomic lines were largely filled in, a comparison was feasible of calculated OH and CH line strengths against those of the Sun and stars. Because solar OH lines were otherwise modeled too strong, all the transition probabilities were reduced by an additional 0.15 dex uniformly. In part because weak OH lines in the Sun were reasonably well modeled but strong ones were still too strong, a chromospheric mitigation of the surface temperature drop was invoked as described shortly. As illustrated below in Figs. 3d and e, these changes suffice to reproduce both strong and weak 0-0 OH features in stars as well as the Sun. -- 12 -- The calculations were first run using Kurucz (1995) models, but the very metal- poor stars showed a Teff deduced from Hα as much as 200 K higher than that from fitting the mid-UV. That discrepancy was removed by adopting Castelli et al. (1997) models, models in which convective overshoot has been turned off. We did not use the alpha-enhanced Castelli et al. (1997) models, as they are available for only a single value of the alpha-element abundance enhancement. Like the Kurucz (1995) models, the Castelli et al. (1997) models we used were calculated with χt = 2 km s−1 and a logarithmic solar iron abundance of −4.37 with respect to the abundance by number of hydrogen plus helium, which amounts to −4.33 with respect to hydrogen alone. These choices are about 1 km s−1 higher in χt than the ∼1 km s−1 typically found for the solar flux spectrum, and about 0.15 dex higher in iron abundance than indicated by the meteoric iron abundance and by several recent determinations for the Sun (e.g., Bi´emont et al. 1991; Holweger, Kock, & Bard 1995). Thus both sets of models were calculated with higher line blanketing than expected for a solar-type star of the nominal model metallicity, since iron lines dominate the UV and optical spectrum, and the higher χt increases the strengths of moderately strong lines of every species. For each star including the Sun, a Castelli et al. (1997) model was interpolated to the nearest 50 K in Teff (25 K for the Sun), 0.1 dex in log g, and 0.05 dex in [Fe/H]. The microturbulent velocity χt was changed at all depths to the value indicated for each star in the table and figures below. To minimize the overblanketing effect, our models were interpolated to a nominal grid abundance 0.10 -- 0.15 dex lower than that desired. To set abundances, the abundance of the interpolated model was then raised to the desired value and the relative abundance of iron dropped by 0.15 dex. The specific light-element abundance elevations required by the optical observation were adopted, or assumed to be zero for the Sun and α Cen A , and all species equilibria were recalculated. This procedure incorporates the specific χt and iron and light-element abundances set for the model in calculating both the ionization and molecular equilibria and the line and continuous opacities for its spectrum. Concerning the model structure, the procedure compensates for the effects of overblanketing due to the high choice of iron abundance, but not for the light-element ratio nor for the microturbulence. We anticipate very mild errors in model colors and temperature structure until spectra become dominated by molecular lines or otherwise are considerably stronger-lined than any of these. The comparison of model and observed colors (§6) bears out this expectation. For the Sun, a model with Teff = 5775 K , log g = 4.4, [Fe/H] = 0, [α/Fe] = 0, and χt = 1.0 km s−1 , was interpolated in the above way. For HD 128620 = α Cen A we adopted a similarly interpolated model with Teff = 5800 K, log g = 4.3, [Fe/H] = +0.15 dex, [α/Fe] = 0, χt = 1.0 km s−1 . While Neuforge-Verheecke & Magain (1997) found Teff = 5830 K and [Fe/H] = 0.25 dex, most other analyses have derived somewhat lower Teff and [Fe/H] values. At Teff = 5800 K and χt = 1.0 km s−1, [Fe/H] = 0.15 dex provides a better match to the mid-UV spectrum of α Cen A. Even with the use of Castelli et al. (1997) models interpolated as above, another major type of mismatch persisted in all strong-lined stars: cores of lines formed near the surface appeared -- 13 -- too strong in the calculations, especially in extremely strong mid-UV lines such as those of Mg ii and Mg i or Fe ii and Fe i. This is definitely not due to wayward gf-values or damping constants, because of the trend seen in the mismatch. Profiles calculated for the strongest lines in the most metal-poor spectra generally matched or were somewhat too high in the core, but were somewhat too low in the core in stars of intermediate line strength, and became too low across the the central ten Angstroms or more in the strongest-lined stars. (See Fig. 3a below for residual trends of this nature.) As noted above, a related mismatch was seen within the OH spectrum itself: in any given strong-lined spectrum, notably that of the Sun for which the model parameters are extremely well known, weak OH lines were matched at a higher oxygen abundance than strong lines. The possibility discussed in §5 that the solar continuum may be drawn too low would lead to a mismatch of opposite sign. To the extent that the gf-values are reliable, then, this indicates an error in the temperature structure of the model. The most obvious possibility is a chromosphere, which is not incorporated in flux-constant, static models such as those of SYNTHE. The solar chromosphere, for example, is seen in images of the limb and evident from emission reversals in the cores of strong spectral lines such as Mg ii. It is generally modeled by an increase in temperature in relatively shallow layers of the atmosphere (e.g., Fontenla et al. 1999). Any of several non-radiative heating sources extending from the photosphere through the chromosphere into the corona could be responsible (e.g., Schrijver 1995; Schrijver et al. 1999; Sterling 2000). We consequently changed the temperature of individual points in each interpolated stellar model with [Fe/H] & −1.0, beginning just below the point where the model becomes optically thin at 5000 A (typically at depth number 51 or 52). (We tried this for the more metal-poor models as well, but the profiles of all the strong lines became much too high near the core, so we left these models as they were.) To best fit both optical and mid-UV spectra simultaneously, we increased the temperature at model depths 28 through 48 by 10, 20, 30, 40, 50, 60, 70, 80, 90, 100, 110, 120, 130, 140, 150, 150, 140, 120, 90, 60, and 30 K respectively. This temperature enhancement may be considered an approximate chromosphere, though it is based entirely on empirical fits of one-dimensional spectral line calculations to observed spectral line profiles. It incorporates neither energy transport considerations nor matches to spatially resolved structures, both of which lie beyond the scope of this work. 5. Matching Mid-UV Spectral Calculations The degree to which our mid-UV and optical spectral calculations match spectra of real stars is illustrated in the figures. The first three show mid-UV spectra, and the last, optical. Within a figure, panels are ordered by wavelength. Flux (central intensity for the solar mid-UV spectrum) is plotted versus wavelength in air. The spectrum calculated for each star (light line) is plotted on top of that observed (heavy line). -- 14 -- In Figure 1, the global mid-UV fit is shown for the nine stars with absolute fluxes. All spectra have been Gaussian smoothed to a FWHM resolution of 500 (550 for HD 128127 and α Cen A ). Each spectrum is displaced vertically by ∼20% from the one below. For each star a single mid-UV continuum normalization constant was adopted, as listed on the left. On the right, the HD number of each star appears in bold. Below are found the model-atmosphere parameters adopted for the corresponding calculation: Teff (in K), C wherever the chromosphere was modified as described at the end of §4, log g, [Fe/H], χt (in km s−1). On the following line appear the values in dex of the light-element abundance enhancements [Element/Fe] (in dex), first that of oxygen, then those for Mg, Si, Ca, and Ti. (The latter four are grouped together wherever they are the same.) While the fits are generally good, serious discrepancies are noted in α Cen A at the bottom of the figure near 2550 -- 2600 A and 2640 -- 2710 A, regions of relatively little line absorption in the more metal-poor spectra. Figure 2 illustrates the effect on mid-UV fits when model parameters are changed somewhat. Two of the stars of Fig. 1 are depicted: the moderately weak-lined HD 94028 (top two spectra) and the rather strong-lined HD 184499 (bottom three). Star and model identifications and normalization constants are indicated as in Figure 1. The lower plot for each star represents its best fit, a replot of that star's comparison in Figure 1. Comparing the bottom plot to the middle plot shows the effect for HD 184499 of decreasing [Mg/Fe] by 0.3 dex to the solar [Mg/Fe] ratio. Although minimized by the adoption of a normal- ization constant 11% higher, the mismatch is still evident. The strengths of the wings of the strong Mg ii doublet at 2800 A (and of Mg i at 2852 A ) are poorly reproduced, as is the overall pseudo- continuum level near and below the bound-free edge of Mg i at 2512 A . Similar calculations run for HD 94028 show a more dramatic effect in the Mg ii and Mg i line wings, which are weak enough that little renormalization is possible, but a smaller effect near 2512 A , where the H− opacity is dominant because of the reduced magnesium abundance. Comparing the plot just above the best fit for each star to the best fit itself shows the degeneracy of the mid-UV spectra with respect to a simultaneous change in Teff , log g and [Fe/H] designed to preserve overall line strength and ionization equilibrium. For each plot just above the best fit, Teff was decreased by 100 K , [Fe/H] and log g both concurrently decreased by 0.1 dex, and the spectrum renormalized accordingly. This change is very difficult to detect from fitting the mid-UV slope alone at these metallicities and this resolution. The increased metallicity reduces the blue pseudocontinuum relative to the red at the same time that the increased Teff raises it. The change is marginally detectable in HD 94028 from the mismatch of low-opacity windows, but not in HD 184499 because the overall match is less satisfactory. Such a change does lead to a reduction in the normalization parameter, by 11% for HD 94028 and 17% for HD 184499. In principle this might be detected by other means, but only if [Mg/Fe] is properly modeled. For as seen above, taking [Mg/Fe] = 0 (as in the Kurucz flux distributions) can largely compensate in stars such as HD 184499 that show a light-element enhancement. -- 15 -- Our conclusion, then, is that if the mid-UV normalization constant is treated as a completely free parameter (as here), the mid-UV flux distribution by itself can currently be used to set Teff only in MSTO stars with [Fe/H] ≤ −1.5. While the mid-UV was useful in demonstrating the need for models with no convective overshoot and with an approximate chromosphere, our determinations of Teff for the more metal-rich MSTO stars have relied on the optical spectra instead. We show in Fig. 4 below that the Balmer line profiles and the strengths of other high-excitation lines are well matched in the optical spectra of all stars at the same Teff as the best-fit mid-UV spectrum, and in §5 that model colors match observations as well, so that Teff is indeed very well established here. However, our goal of Teff determination for turnoff stars of all metallicities from mid-UV spectra alone is not yet met. We are not able to calculate the low-opacity windows blueward of 2900 A accurately enough at near-solar metallicity, as we now demonstrate. Figure 3 plots at higher resolution and higher scale the same mid-UV spectral comparisons as are shown in Figure 1, plus that of the Sun. Panel a shows 2627.5 -- 2630.5 A, dominated by strong lines. The other panels show regions of locally low line blanketing: b, 2642.5 -- 2645.5 A; c, 2900.0 -- 2903.0 A ; d, 3082.5 -- 3085.5 A ; e, 3100.0 -- 3103.0 A . In all plots in Fig. 3, the spectrum of HD 184499 is offset by 30% from that of HD 128620 = α Cen A, which in turn is offset by 50% from that of the Sun, for which 1.0 marks the height of the true continuum. The same normalization constant as in Fig. 1 has been adopted for each stellar spectrum, as was a normalization for the Sun 1.10 times that given in the Kurucz & Avrett (1981) atlas, as noted on the left. Identifications are provided at the top for the strongest lines as calculated in the solar spectrum. Manually added lines are flagged with a colon following the decimal digits of the wavelength in Angstroms. Next appears the species, the lower excitation potential in eV of atomic transitions or the band and line of molecular ones, the digits following the decimal of the residual flux in the unbroadened line of the solar spectrum calculation, and the log gf-value adopted. In Fig. 3e of the 3100.0 -- 3103.0 A region, calculated and observed spectra agree reasonably well in all stars, including α Cen A and the Sun. No "missing" lines are strong enough to merit explicit identification at the top of the figure. True continuum is reached in each spectrum, but fleetingly, near 3102.5 A ; its placement is sensitive to the treatment of adjacent weak lines. An unblended 0-0 OH line appears at 3101.230 A , next to a near- continuum high point; both are modeled well in all spectra. The strong atomic lines are well reproduced, as are two moderately strong CH lines at 3100.120 and 3100.170 A. Fig. 3d offers a further picture of how well both continuum and OH features are modeled. Another continuum window appears near 3084.7 A , but again with a weak line superimposed in the solar and α Cen A spectra. The relatively unblended OH 0-0 band line at 3084.896 A , used in several oxygen-abundance investigations, is well reproduced over the factor-of-100 metallicity range of the stars in whose spectra it is detected, including the Sun itself. Nearby at 3085.2 A is a blend dominated by a stronger 0-0 OH line; it too is reasonably well matched in all stars. Recall that to achieve this agreement, the Gillis et al. (2001) theoretical gf-values for OH were reduced by 0.15 dex; otherwise these features and that of Fig. 2e, plus the overwhelming majority of other -- 16 -- OH lines, would be too strong in the Sun. That is, unless the solar continuum were raised. The solar continuum appears to be well defined, and the OH lines of all bands matched to ∼0.1 dex, throughout the 3000 -- 3100 A region. Because there are so few continuum windows, however, we cannot rule out a systematic underestimate of the continuum in α Cen A and the Sun of perhaps 5% or 10%. Roughly speaking, raising the α Cen A mid-UV continuum would allow it to be modeled at an [Fe/H] value more consistent with that of Neuforge-Verheecke & Magain (1997), while raising the solar continuum might eliminate the present need to reduce the theoretical OH gf-values by 0.15 dex. As noted at the end of §6, it would also bring the mid-UV normalization constants of the Sun and α Cen A closer to their expected values. A definite answer could be obtained by deriving the stellar parameters of α Cen A from an optical spectrum in the same way as for the other stars, and from modeling an improved spectrum of the solar near-UV. Random continuum variations approaching 5% occur for two, and possibly all three, of the stars that were noted in §3 as observed through a smaller aperture. For the continuum seems appropriately placed for all stars in Fig. 3e, but in Fig. 3d the continuum of the HD 106516 observed spectrum appears too low, and in Fig. 3c that of HD 128167 may also be too low. Such variations might also be present in the HD 128620 = α Cen A spectrum as well, but remain undetected because of the difficulty of matching its multitude of lines. Fig. 3c shows another low-opacity window at 2900 A. Here the number of missing lines added remains small, and agreement is nearly as good. OH lines are less important, being generally minor constituents of blends. However, windows of true continuum have disappeared from the solar and α Cen A spectra, and are fleeting at best in the echelle spectra of lower resolution that were acquired for mildly metal-poor stars. Continuum placement depends not so much on OH features but rather on whether all missing lines have been spotted. Clearly that is not the case at solar metallicity, where weak missing lines apparently overlap to the extent that they cannot be distinguished as individuals. Although the cumulative effect of their neglect on the total absorption is small, is it significant for the continuum placement, which consequently cannot be determined as well as was the case above from fits such as this. It is placed here and for the remaining panels by adopting the same normalization factors required to match Figs. 3d and e. By 2645 A , in Fig. 3b, the fits have deteriorated considerably. This is especially true at and above solar metallicity, where the numbers of missing and identified lines noted at the top are comparable. True continuum still appears in the weakest-lined spectra near the top of the panel, but elsewhere it is gone, lost to a forest of weak lines. Although the fit to the metal-poor stellar spectra is still reasonable, the high-resolution spectra of the α Cen A and the Sun reveal that too few missing lines have been added, often too strong and in the wrong places. For the adjacent region from 2645 to 2700 A, the sheer number of missing lines precluded their reliable identification at the low resolution of the low-metallicity stellar spectra. Because many were simply omitted, the calculations overestimate the spectral flux in that region -- mildly in moderately metal-poor stars but seriously at and above solar metallicity, according to Fig. 1. -- 17 -- From Fig. 3a, it is clear that these difficulties are reduced in regions dominated by strong atomic lines. There the fit is dictated more by the ability of the calculations to match the cores and the damping wings of strong lines, and not by modeling the weak or missing features. These are largely suppressed: although the region shown is blueward of that of panel b, only one missing line is noted in it. The wings of the strong lines are reproduced well in the four top spectra but begin to be somewhat too broad near the core in HD 106516 and HD 201891. The entire ±1 A about the core is too deep in α Cen A and the Sun. This could perhaps be remedied by continuing the temperature enhancement to shallower surface layers than those indicated in §4. Figure 4 shows fits to the optical spectra of the Sun and all stars except α Cen A. Identifications are given for the strongest lines in the solar spectrum. Since the optical spectra were not fluxed, the continuum was normalized to unity. The same models were used to calculate the synthetic spectra as in Figures 1 and 2. The wavelength regions in panels a -- e depict a variety of magnesium lines and those of other heavier elements, to illustrate the extent to which the spectral calculations match features of a wide range of excitation, ionization, and strength. Figure 4f shows Hα, to illustrate the satisfactory choice of Teff . The normalization of the Hα profile was aided by its position near the center of an echelle order, and was done by interpolating the normalization functions of the two adjacent orders, centered ±75 A away. This was straightforward, for the Hα line is much narrower than this and the continuum is very well defined by a spline fit of low order; moreover, the normalization functions for the adjacent orders were very similar. Note that most absorption lines in this panel are telluric, not stellar, as evidenced by their shift in apparent wavelength from one star to the next. Included in Fig. 4f above the best fit for HD 94028 is the spectrum for the cooler HD 94028 model whose mid-UV spectrum is shown in Fig. 2. This illustrates how the fit to Hα deteriorates when models of 100 K lower Teff are adopted. The Hα profile is relatively insensitive to gravity in this case: calculations adopting a model that differed from the best-fit model only in having log g 0.2 dex larger produced a profile that deviated only near 6560 A and 6565 A , by about the thickness of the heavy line. Similar calculations showed no sensitivity of the Hα profile to the chromospheric elevation of temperature, nor to changes of 0.3 dex in iron abundance or light-element ratio. However, Kurucz (1995) models required a Teff 200 K hotter, thanks to their inclusion of convective overshoot. Panels a -- e of Fig. 4 show that all the optical magnesium features are well matched using the same magnesium abundance and the same model as for the best fit to the mid-UV and to Hα. Panel a plots the region near the very high-excitation Mg ii doublet at 4481.2 A; b, the ground-state Mg i line at 4571.1 A; c and d, the Mgb lines at 5167.3, 5172.7, and 5183.6 A; and e, the high-excitation Mg i line at 5711.1 A . These panels also demonstrate that the calculations match strengths and profiles of other atomic lines regardless of species, excitation, or strength. Among the lines depicted are Ti i at 5173.7 A, Ti ii at 5185.9 A, ground-state Fe i at 5166.3 A , Fe ii at 4576.3 A , and Si i at 5708.4 A. These panels also contain missing features: not modeled at all in the calculations are the -- 18 -- absorption lines seen near 4480.8 A, 5170.7 A, and 5181.3 A. They are not telluric, since they appear at the same position in the stellar spectra, and are especially visible in the two stronger-lined stars. 6. Results and Comparisons Our results for the parameters of the eight metal-poor turnoff stars analyzed here are listed in columns 12 -- 15 of Table 1. The Teff values found for these stars by works cited in §1 are included for reference. Our Teff values are seen to fall in the middle of the range of previous results, preferring neither the hotter nor the cooler values. Unfortunately, the number of stars is too small and the scatter too large for more definitive conclusions. The reliability of our Teff values is supported by a comparison of model colors to those observed. For each star including the Sun, we interpolated the U BV and b − y colors and V magnitudes calculated by Castelli et al. (1997) to the Teff , log g, and [Fe/H] values of each model we used. We have made no reddening corrections, nor any corrections for the temperature enhancement invoked in the more metal-rich models as described at the end of §4. The ninth, tenth, and eleventh columns of Table 1 show the difference in B − V , U − B , and b − y between the observed and calculated colors for each star. The agreement is extremely good in all colors over the entire metallicity range. The mean difference is −0.010, −0.018, and 0.000 mag in B − V , U − B , and b − y respectively. The standard deviation of an individual difference is 0.010, 0.026, and 0.015 mag, marginally larger than typical observational errors alone. The color differences are small even for α Cen A, for which both metallicity and Teff are influential (we interpolated to a model [Fe/H] = +0.05), and for which we have no optical spectrum and so did not derive parameters. For the other stars, B − V and b − y are sensitive primarily to Teff ; their deviations correspond to ±45 K with no allowance for observational error. Their observed colors are slightly bluer than those of models, but the size of the differences points to a scale error in Teff of less than 50 K. We thus estimate an uncertainty of ±50 K in each Teff determination, from this and the poorer quality evident in Figs. 2 and 4f of the fits to mid-UV fluxes and Hα profiles as Teff is increased by 100 K, plus the deterioration of the fit shown in Fig. 4a to Mg ii. From plots like those of Fig. 4 at other wavelengths, we estimate the uncertainty in [Fe/H] to be ±0.05 dex at fixed Teff , increasing to ±0.1 dex when the Teff uncertainty is included. The mid-UV slopes and Hα profiles are only weakly sensitive to log g, so our results for log g are susceptible (as are those of other groups relying on ionization equilibria) to errors in gf-value scales and to possible non-LTE effects. We have not yet taken full advantage of the technique of matching damping wings of strong lines, incorporating improved treatments of hydrogen broadening such as those of Anstee & O'Mara (1995) and Barklem & O'Mara (1997). It would be valuable to do so given the constraints this would offer on alternative measurements of the fundamental properties of these standard stars, for example parallaxes from HIPPARCOS and other space missions. However, the following comparisons of observed and model fluxes indicate that the log g values are good to ±0.1 dex as they stand. -- 19 -- We first compare the UV normalization constant with that expected. According to Mihalas & Binney (1981), the ratio of observed to emitted stellar flux is f /F = θ2/4, where the angular diameter θ = 2R/D, R being the stellar radius and D its distance. For each star we calculate log R in solar units as 0.5(log M − log g + 4.4377), assuming masses M = 0.8 M⊙ for for the metal- poor stars based on stellar isochrones (VandenBerg et al. 2000) and 1.15 M⊙ for HD 128167, and adopting 1.16 M⊙ for α Cen A as derived from its orbit by Pourbaix et al. (1999). The distance in solar units is found from the parallax, and the theoretical flux normalization constant follows from F/f times 4.2545 ×1010 to convert to square steradians. The ratio of this theoretical constant to that actually used to scale the observations in Figs. 1 and 3 is listed in column 7 of Table 1. Agreement is good for all stars but one. For α Cen A, the ratio is 0.46, suggesting an observed flux that is too low. Indeed, Linsky & Wood (1996) saw this in their 1995 mid-UV α Cen A HST observations with GHRS: relative to earlier GHRS and IUE observations of α Cen A, "our fluxes are lower (by factors of 5.6 and 4.8 for the Mg ii and Fe ii spectra, respectively), presumably because of the star's being misplaced at the edge of the small aperture." From a visual comparison with Fig. 1 of Linsky & Wood (1996), the fluxes we are using for the Linsky et al. (2000) STIS α Cen A spectrum should be raised by a factor of 2.1. This brings the α Cen A ratio to 0.96, and results in an average of the nine stellar ratios of 1.06 ± 0.05, with an individual standard deviation of 0.15. For the Sun, the ratio is 0.91 given the scale factor of 1.10 noted above, which is within the stated normalization uncertainty but still low. These α Cen A and solar ratios would be closer to unity were the continuum levels for their spectra drawn higher by 5% -- 10%, a possibility raised in §5. As matters stand, errors in continuous opacity are limited to <10% nonetheless. The apparent V magnitude may also be compared to that expected from the model. From the V magnitudes tabulated with the model colors, we calculated the difference in the stellar model V with respect to the solar model, added this and 5 log(D/R) to the observed solar V = −26.7, then subtracted this from the observed stellar V magnitude. Again, no reddening or other corrections were applied. The resulting differences are shown in column 8 of Table 1. The agreement is reasonable: the theoretical magnitudes average 0.09 ± 0.06 mag brighter than those observed, with a ± 0.18 mag individual standard deviation. 7. Summary and Discussion In brief, we have calculated high-resolution mid-UV spectra from first principles, after modify- ing line parameters and assigning line identifications for missing lines based on changes seen among spectra of stars spanning a range of well-determined values of Teff , log g, and [Fe/H]. The resulting calculations provide a good fit for turnoff stars of all metallicities redward of 2900 A . Previous discrepancies in fitting the solar pseudocontinuum (noted in §1) are attributed not to missing con- tinuous opacity, which is good to <10% near 3100 A , but rather to missing or miscalculated line opacity. -- 20 -- To improve the accuracy of the mid-UV gf-values, damping constants, and line identifications derived here, we have determined temperatures, gravities, microturbulence, and abundances for eight stars by calculating fits to optical as well as UV spectra. For stars more metal-poor than one-thirtieth solar, the fit to the mid-UV spectrum was satisfactory using laboratory lines alone, and can by itself determine Teff . For the more metal-rich stars, lines missing from the laboratory list compromise the goodness of fit of the mid-UV spectrum. A 100 K increase in Teff is seen to be difficult to detect from mid-UV spectra alone, as long as [Fe/H] and log g are increased by 0.1 dex to compensate. This Teff -[Fe/H] degeneracy parallels the age-metallicity degeneracy in elliptical galaxies (Worthey 1994, 1999). As with galaxies, it is reduced here by considering the Balmer lines, with the advantage that stellar models need not include Balmer-line emission nor composite populations. Here Teff was set for stars with [Fe/H] ≥ −1.5 by matching the Hα line profile and the strengths of other high-excitation features. For all the metal-poor stars, the excellent match between calculations and observations of both optical and mid-UV spectra, and between model and observed colors, indicates an uncertainty of ±50 K in an individual Teff determination, and a similar uncertainty in the Teff scale. Moreover, the stellar V magnitudes generally agree to 0.1 mag, and the mid-UV normalization constants to 10%, with those expected from the stellar angular diameter found from the observed parallax and the model log g, assuming reasonable masses. In this process, it was found that these optical diagnostics and the slope of the mid-UV con- tinuum could only be fit simultaneously when using models of Castelli et al. (1997) in which convective overshoot has been turned off. Our work thus agrees with that of Castelli et al. (1997), who found that their models better reproduce the Balmer profiles in the Sun than do those of Kurucz (1995). Downloading the Castelli et al. (1997) models from the Kurucz web site at http://cfaku5.harvard.edu is recommended for all work where the results depend upon the tem- perature structure in deep continuum-forming regions. Residual discrepancies in fitting the profiles of very strong atomic lines, and the strengths of strong versus weak solar OH lines in the 3000 -- 3100 A region, are reduced here by mimicking a chromosphere towards the surface of every model with [Fe/H] & −1. As described at the end of §4, the model temperature was raised in the region just above the depths where optical continuum and weak lines are formed. While this procedure was necessary to match strong-line cores and OH line strengths over a very broad range of metallicity with a single set of gf-values, it is definitely an oversimplification. Energy and stability considerations remain to be examined. Fits might be improved with a different choice of temperature enhancement. Indeed, a two-stream model might be found necessary. In the meantime, line cores also will be subject to uncertainties due to the presence of chromospheric emission. Peterson & Schrijver (1997) demonstrate that Mg ii emission is always present in solar-temperature stars, with a strength that increases with increasing metallicity. Emission also increases with increasing stellar activity (Ayers et al. 1995). Since activity in turn is associated with rapid rotation, stronger Mg ii emission appears commonly among young stars and close binaries. Consequently, no single quiescent chromosphere appears capable of modeling the -- 21 -- very core of the profile of a strong line in all turnoff stars. The reliability of our Teff values and the good match achieved for OH lines enable a second look at the oxygen abundances of metal-poor stars, which are still uncertain as noted in §1. Our values for Teff are intermediate, suggesting that Teff is not the sole cause of the oxygen discrepancies. We defer a more detailed discussion to a future paper by Peterson, Dorman, & Rood (2001), who compare oxygen abundances derived from the OH lines with those from the permitted and forbidden lines of atomic O i in these stars. The manual addition of lines to the model line list, to match otherwise missing features seen in the observed spectrum, is clearly necessary for even an approximate representation of the mid- UV spectrum of solar-temperature stars of one-tenth solar metallicity and higher. Nonetheless the procedure has its important limitations. One is that as overall line strength increases, the proportion of missing lines goes up dramatically; disentangling them requires a broader range of spectra of high quality, as discussed below. Another is that the excitation or ionization of the transition could be misjudged; the ramifications are discussed in the next paragraph. A third is that many missing lines may not be due to iron as assumed, but to other elements. As discussed in §4, this is potentially serious redward of the bound-free absorption edge of Mg i near 2512 A. Elsewhere the effect should be minimal, because although other iron-peak elements contribute significant line absorption throughout the mid-UV, the relative abundances of all other iron-peak elements with respect to iron change little from one late F or early G main-sequence star to another in both halo and disk populations in our Galaxy. To be sure, the list generated by this procedure cannot be expected to reproduce the spectrum of a peculiar A or F star, for example a Cr-enhanced star, in which the iron-peak ratio has been dramatically altered (Jaschek & Jaschek 1990). More generally, the list is not suitable for modeling stars whose parameters lie well outside the range of those analyzed here. In such cases, any difference becomes important between the assigned and the actual excitation (and ionization) potential of the transition responsible for a line, leading to a significant misrepresentation of its strength. Consequently, the present list cannot be expected to reproduce the spectra of stars hotter than early F nor cooler than early G, nor giants, as well as stars of metallicity substantially higher than solar for which lines are still missing. The list could be made suitable for those stellar types by extending the present analysis to include such stars, provided high-resolution, high S/N mid-UV spectra are available in which mismatches induced by errors in the list can be visually identified and corrected. We discuss this further below. Conversely, the list as it stands should do as well as indicated by Fig. 1 and Fig. 3 in reproducing the mid-UV spectra of any other late F or early G star of solar metallicity or less, provided a suitable model atmosphere is chosen. This holds because once the gf-value and damping constant of a line transition are set, elemental abundance and the Boltzmann and Saha equilibria at each point in the atmosphere fully determine line strength in quiescent stars such as these. Dynamical effects can be ignored, and non-equilibrium effects (if any) are slowly-varying functions of spectral type, hence absorbed into the gf-value determinations. The only exception is the chromospheric emission noted -- 22 -- above. Were the line list improved to the point that a reliable match is achieved to the mid-UV spectra of F and G stars of solar metallicity and above, we could expect to use the mid-UV slope to at least constrain age and metallicity, regardless of its overall level or the appropriate light-element abundance ratio, as these could be incorporated in a grid of calculations covering the entire relevant parameter space. Spectral indexes offer an alternative means by which mid-UV spectral calculations alone might constrain metallicity and Teff of a star or age of a stellar system. The principle is to construct indexes based on spectral regions dominated by features whose behavior breaks the age/metallicity degeneracy and elucidates the alpha enhancement, examples of which are the Balmer lines and ratios of Fe i/Fe ii and Mg i/Mg ii line strengths. Spectral indexes have indeed been used as diagnostics of age and metallicity in ellipticals by several of the groups mentioned in §1. For example, Lotz, Ferguson, & Bohlin (2000) have compared mid-UV and optical line indexes they calculated from Kurucz flux distributions against mean values observed for the stars considered by Fanelli et al. (1990). Unfortunately, poor agreement was seen for all but two mid-UV indexes, the slope of the pseudocontinuum between 2600 A and 3100 A, and blended absorption near 2538 A. We should be able to improve upon some of those index calculations immediately. As Figs. 1 and 3 illustrate, the current spectral calculations should be reasonably reliable at all metallicities in spectral regions redward of 2900 A , and for those dominated by strong lines blueward of that. This includes several indexes already in use, notably the Mg ii and Mg i lines at 2800 A and 2852 A and line blends in the 3000+ A region. We plan to begin such index calculations shortly, and make them generally available. Blueward of 2900 A, however, our calculations are not yet reliable enough. Fig. 1 shows that our calculated spectra of solar-metallicity stars substantially overestimate the true fluxes in those wavelength regions not dominated by strong lines. Somewhat ironically, our calculations thus suggest that neither of well-behaved Lotz, Ferguson, & Bohlin (2000) indexes is currently reliable at near-solar metallicity. The problem undoubtedly is lines still missing from the line list. As seen in Fig. 3, missing lines often overlap in the solar spectrum to such an extent that they could not be individually discerned, and so their species and wavelength assignments became arbitrary. This was true especially in the 2550 -- 2600 A and the 2640 -- 2710 A regions, where we did not complete the process. This is the reason for the excess flux seen in the calculations at solar metallicity in these regions. This could be remedied were spectra obtained of the same quality as that of α Cen A for stars of the line strength of HD 184499 that span a range of properties, notably Teff . For example, the spectra of HD 184499 and a Hyades F star with Teff ∼6500 K should have about the same overall line strength: the Hyades star's higher abundance, [Fe/H] ∼+0.15, is offset by its higher Teff at the rate of 0.1 dex per 100 K, as Fig. 2 illustrates. Because line crowding is much lower than in α Cen A , comparing high-resolution mid-UV spectra of HD 184499 and a sharp-lined Hyades F star could show where the strongest lines are that are missing at solar metallicity and suggest their identities -- 23 -- from the changes observed with Teff . The α Cen A and solar spectra could then be used to fill in the rest. Analyzing an additional star closer in Teff to HD 184499 but with [α/Fe] = 0 would show explicitly which transitions are due to magnesium in the 2550 A region, and whether OH has been properly characterized throughout the 2900 -- 3100 A domain. At that point, the full arsenal of indexes representing the strengths of mid-UV features could be calculated from the spectra, and their behavior with Teff , log g , [Fe/H], and [α/Fe] assessed explicitly. This is a major advantage of calculations from first principles over empirical libraries, which are confined to the spectra of stars available nearby. It is a major goal of this project. Indexes should result with significant diagnostic power for the integrated-light spectrum of an old stellar system. We are indebted to S. Allen of U. C. Santa Cruz for writing the initial scripts to convert the VAX version of Kurucz codes into UNIX versions. We are grateful to M. Bessell of Mount Stromlo and Siding Spring Observatories for communicating the Gillis et al. (2001) and LIFBASE gf-value calculations and programs for their comparison against the Kurucz values, to J. Valenti of the Space Telescope Science Institute for providing the HD 128620 spectral reductions, to T. Misch of Lick Observatory for obtaining the blue spectrum of HD 114762, and to J. Fulbright of the Dominion Astrophysical Observatory for contributing his optical echelle spectra. We thank Fulbright and R. P. Kraft of Lick Observatory, R. L. Kurucz of the Harvard-Smithsonian Center for Astrophysics, T. M. Lanz and S. R. Heap of the NASA/Goddard Space Flight Center, D. C. Morton of the Herzberg Institute for Astrophysics, and R. W. O'Connell of the University of Virginia for useful discussions. This work has made use of the SIMBAD database, operated at CDS, Strasbourg, France. Support for this work was provided by NASA Astrophysics Data Program contract S- 92512-Z, STScI grants GO-07395, GO-07402, and AR-8371, and NSF grants AST 99-00582 and AST-0098725, to Astrophysical Advances, and by NASA ADP grant NAG 5-7104, NASA LTSA grant NAG 5-6403, and STScI grant GO-06607 to the University of Virginia. REFERENCES Allende Prieto, C., & Lambert, D. L. 2000, AJ, 119, 2445 Allen, C. W. 1973, Astrophysical Quantities, 3rd. ed. (London: Athlone), 162 Alonso, A., Arribas, S., & Martinez-Roger, C. 1996, A&A, 313, 873 Anstee, S. D., & O'Mara, B. J. 1995, MNRAS, 276, 859 Ayres, T. D., et al. 1995, ApJS, 96, 223 Balachandran, S. C. 2000, in Proc. IAU JD8, in press Balachandran, S., & Bell, R. 1998, Nature, 392, 791 -- 24 -- Barklem, P. S., & O'Mara, B. J. 1997, MNRAS, 290, 102 Bi´emont, E., Badoux, M., Kurucz, R. L., Ansbacher, W., & Pinnington, E. H. 1991, A&A, 249, 539 Blackwell, D. E., Petford, A. D., Shallis, M. J., & Simmons, G. J. 1982, MNRAS, 199, 43 Blackwell, D. E., & Lynas-Gray, A. E. 1998, A&AS, 129, 505 Boesgaard, A. M., King, J. R., Deliyannis, C. P., & Vogt, S. S. 1999, AJ, 117, 492 Butler, et al. 1993, J. Phys. B, 26, 4409 Carney, B. W., Latham, D. W., Laird, J. B., & Aguilar, L. A. 1994, AJ, 107, 2240 Castelli, F., Gratton, R. G., & Kurucz, R. L. 1997, A&A, 318, 841 Chmielewski, Y., Friel, E., Cayrel de Strobel, G., & Bentolila, C. 1992, A&A, 263, 219 Dorman, B., Rood, R. T., & O'Connell, R. W. 1993, ApJ, 419, 596 Fanelli, M. N., O'Connell, R. W., Burstein, D., & Wu, C.-C. 1990, ApJ, 364, 272 Fontenla, J., White, O. R., Fox, P. A., Avrett, E. H., & Kurucz, R. L. 1999, ApJ, 518, 480 Fuhr, J. R., Martin, G. A., & Wiese, W. L. 1988, J. Phys. Chem. Ref. Data, 17, Suppl. 4 Fuhrmann, K., Axer, M., & Gehren, T. 1994, A&A, 282, 585 Fulbright, J. P. 2000, AJ, 120, 1841 Furenlid, I., & Meylan, T. 1990, ApJ, 350, 827 Gillis, J. R., Goldman, A., Stark, G., & Rinsland, C. P. 2001, J. Quan. Sp. Rad. Transfer, 68, 225 Gratton, R. G., Carretta, E., & Castelli, F. 1996, A&A, 314, 191 Heap, S. R., et al. 1998, ApJ, 492, L131 Henry, R. B. C., & Worthey, G. 1999, PASP, 111, 919 Holweger, H., Kock, M., & Bard, A. 1995, A&A, 296, 233 Israelian, G., Rebolo, R., Garc´ıa L´opez, R. J., Bonifacio, P., Molaro, P., Basri, G., & Shchukina, N. 2000, ApJ, 551, 833 Jaschek, C., & Jaschek, M. 1990, The Classification of Stars (Cambridge: Cambridge Univ. Press) King, J. R. 1993, AJ, 106, 1206 -- 25 -- Kraft, R. P. 2000, in Proc. IAU 2000, JD8, in press Kurucz, R. L. 1994, CD-ROM 22, Atomic Data for Fe and Ni (Cambridge: Smithsonian Astrophys. Obs.) Kurucz, R. L. 1995, in A.S.P. Conf. Ser. 81, Laboratory and Astronomical High Resolution Spectra, ed. A. J. Sauval, R. Blomme, & N. Grevesse (San Francisco: ASP), 583 Kurucz, R. L., & Avrett, E. H. 1981, "Solar spectrum synthesis. I. A sample atlas from 224 to 300 nm", S.A.O. Spec. Rpt., 391 Kurucz, R. L., & Bell, B. 1995, Kurucz CD-ROM No. 23 (Cambridge, MA: Smithsonian Astro- physical Observatory) Kurucz, R. L., Furenlid, I., Brault, J., & Testerman, L. 1984, Solar Flux Atlas from 296 to 1300 mn (Sunspot, NM: National Solar Obs.), available from http://cfaku5.harvard.edu Lambert, D. L. 2000, in Proc. IAU 2000, JD8, in press Lejeune, T., Cuisinier, F., & Buser, R. 1997, A&AS, 125, 229 Linsky, J. L., Pagano, I., Valenti, J., & Gagne, M. 2000, in Cool Stars, Stellar Systems and the Sun: Eleventh Cambridge Workshop, ed. Garc´ıa L´opez, R. J., Rebolo, R., & Zapatero Osorio, M., A. S. P. Conf. Ser., 223, in press Linsky, J. L., & Wood, B. E. 1996, ApJ, 463, 254 Lombardi, G. G., Smith, P. L., & Parkinson, W. H. 1981, Phys. Rev. A, 24, 326 Lotz, J. M., Ferguson, H. C., & Bohlin, R. C. 2000, ApJ, 532, 830 Luque, J., & Crosley, D. R. 1999, App. Opt., 38, 1423, available from http://www.sri.com/CEM/LIFBASE Magain, P. 1984, A&A, 132, 208 McWilliam, A., Preston, G. W., Sneden, C. & Searle, L. 1995, AJ, 109, 2757 Mihalas, D., & Binney, J. 1981, Galactic Astronomy, 2nd. ed. (San Francisco: W. H. Freeman), 93 -- 96 Morton, D. C. 1991, ApJS, 77, 119 Nave, G., Johansson, S., Axner, O., Ljungberg, P., Malmsten, Y., & Baschek, B. 1994, Physica Scripta, 49, 581 Neuforge-Verheecke, C., & Magain, P. 1997, A&A, 328, 261 Nissen, P. E., Chen, Y. Q., Schuster, W. J., & Zhao, G. 2000, A&A, 343, 722 O'Brian, T. R., Wickliffe, M. E., Lawler, J. E., Whaling, W., & Brault, J. W. 1991, J. Opt. Soc. -- 26 -- Amer. B, 8, 1185 O'Connell, R. W. 1976, ApJ, 206, 370 Peterson, R. C. 1976, ApJ, 210, L123 Peterson, R. C. 2000, in Proc. IAU 2000, JD8, in press Peterson, R. C., Dalle Ore, C. M., & Kurucz, R. L. 1993, ApJ, 404, 333 Peterson, R. C., Dorman, B., & Rood, R. T. 2001, in preparation Peterson, R. C., & Schrijver, C. J. 1997, ApJ, 480, L47 Peterson, R. C., & Schrijver, C. J. 2000, Cool Stars, Stellar Systems and the Sun: Eleventh Cam- bridge Workshop, ed. Garc´ıa L´opez, R. J., Rebolo, R., & Zapatero Osorio, M., A. S. P. Conf. Ser., 223, in press Ponder, J. M., et al. 1998, AJ, 116, 2297 Pourbaix, D., Neuforge-Verheecke, C., & Noels, A. 1999, A&A, 344, 172 Prochaska, J. X., & McWilliam, A. 2000, ApJ, 537, L57 Reid, I. N. 1998, AJ, 115, 204 Schrijver, C. J. 1995, Astronomy and Astrophysics Review, 6, 181 Schrijver, C. J., et al. 2000, Solar Phys., 187, 261 Spinrad, H., Dey, A., Stern, D., Dunlop, J., Peacock, J., Jimenez, R., & Windhorst, R. 1997, ApJ, 484, 581 Spite, M., Francois, P., Nissen, P. E., & Spite, F. 1996, A&A, 307, 172 Sterling, A. C. 2000, Solar Phys., 196, 79 Th´evenin, F., & Idiart, T. P. 1999, ApJ, 521, 753 VandenBerg, D. A., & Bell, R. A. 2000, Proc. IAU 2000, JD8, in press. VandenBerg, D. A., Swenson, F. J., Rogers, F. J., Iglesias, C. A. & Alexander, D. R. 2000, ApJ, 532, 430 Vogt, S. S. 1987, PASP, 99, 1214 Wheeler, J. C, Sneden, C., & Truran, J. W., Jr. 1989, ARA&A, 27, 279 Worthey, G. 1994, ApJS, 95, 107 -- 27 -- Worthey, G. 1999, A.S.P. Conf. Ser., 192, 283 Worthey, G., Faber, S. M., & Gonz´alez, J. J. 1992, ApJ, 398, 69 This preprint was prepared with the AAS LATEX macros v5.0. -- 28 -- Fig. 1. -- Plots are shown comparing observed (heavy line) and calculated (light line) spectra for the nine stars with fluxed mid-UV spectra included in this study. To the right, the name of each star in bold is accompanied by the model parameters used for the corresponding calculation. At the left is the mid-UV normalization constant for each comparison. Fig. 2. -- Plots as in Fig. 1 are shown comparing the fits for two stars with small changes in model parameters. Fig. 3. -- The same as Fig. 1, but on a larger scale in 3 A spectral regions. At the bottom is the Sun: two scans of the center of the solar disk are superimposed. At the very top, the strongest lines in the solar spectrum calculation are identified as described in the text. The same mid-UV normalization constants as in Fig. 1 are adopted for the top nine stars. a) 2627.5 -- 2630.5 A. b) 2642.5 -- 2645.5 A . c) 2900 - - 2903 A. d) 3082.5 -- 3085.5 A. e) 3100 -- 3103 A. Fig. 4. -- Comparisons as in Fig. 1 are shown for optical spectral regions containing magnesium or Balmer lines in eight of the stars plus the Sun. Each observed spectrum (heavy line) has been normalized to unity as described in the text. Line identifications as described for Fig. 3 are shown for the strongest lines in the solar calculation. a) Mg ii. b) Low-excitation Mg i. c), d) The Mgb lines. e) High-excitation Mg i. f ) Hα. In this panel only, the third plot from the top represents a non-optimal choice of model parameters. It illustrates the degradation of the fit to Hα when a Teff is chosen 100 K cooler than that of the plot immediately below. Table 1. Stellar Observations, Model Parameters, and Effective Temperatures HD ⇐ V Observed B − V U − B b − y ⇒ Mid-UV ⇐ V Ratio π Observed − Model B − V U − B ⇒ ⇐ Model b − y Teff log g [Fe/H] ⇒ K93 C+94 A+96 G+96 BL98 F00 χt Teff Teff Teff Teff Teff Teff 19445 +8.06 84937 +8.33 94028 +8.22 106516 +6.11 114762 +7.31 128167 +4.46 128620 +0.00 184499 +6.62 201891 +7.37 Sun −26.7 0.45 0.39 0.47 0.45 0.53 0.36 0.67 0.58 0.51 0.65 −0.25 −0.22 −0.18 −0.14 −0.06 −0.08 +0.23 +0.00 −0.17 +0.13 0.349 0.303 0.343 0.317 0.365 0.254 0.438 0.390 0.358 . . . 25.85 12.44 19.23 44.34 24.65 64.66 742.2 31.29 28.26 . . . 1.37 1.21 1.07 0.99 0.96 0.92 0.46 1.11 0.92 0.91 −0.04 −0.01 −0.01 −0.005 +0.09 −0.01 −0.01 −0.011 +0.25 +0.00 +0.00 −0.005 −0.09 −0.01 −0.03 −0.004 +0.30 −0.02 +0.01 −0.008 −0.05 +0.01 −0.01 +0.008 −0.06 +0.00 +0.01 +0.037 +0.02 −0.02 −0.02 +0.001 +0.40 −0.02 −0.07 −0.008 . . . −0.01 −0.04 . . . 6050 6300 6050 6250 5850 6850 5800 5750 5900 5775 4.5 4.0 4.2 4.3 4.0 4.3 4.3 4.0 4.1 4.44 −2.00 −2.20 −1.40 −0.65 −0.90 −0.35 +0.15 −0.60 −1.00 +0.00 1.0 1.5 1.3 1.0 1.0 2.0 1.0 1.2 1.0 1.0 6007 6314 6047 6221 . . . . . . . . . 5734 6014 . . . 5842 6220 5898 6118 5790 . . . . . . 5726 5817 . . . 6050 . . . . . . 6208 . . . 6707 . . . 5750 . . . . . . 6066 6350 6060 6267 5941 6734 . . . 5750 5974 . . . . . . 6202 5998 6233 5904 6737 . . . 5773 5918 . . . 5825 6375 5900 6200 5800 . . . . . . 5700 5825 . . . Note. -- Units: π in 10−3 ′′; Teff in K; χt in km s−1. Sources: For all stars but HD 128620 = α Cen A, V , B − V , and U − B from Carney et al. (1994); b − y from King (1993), except b − y for HD 114762, HD 128167, and HD 201891 from SIMBAD; π from SIMBAD. HD 128620 = α Cen A: V , U − B , b − y , and π from SIMBAD; B − V from Furenlid & Meylan (1990). All solar observations from Allen (1973). Other Teff sources: K93, King (1993); C+94, Carney et al. (1994); A+96, Alonso, Arribas, & Martinez-Roger (1996); G+96, Gratton, Carretta, & Castelli (1996); BL98, Blackwell & Lynas-Gray (1998); F00, Fulbright (2000). This figure "fig1.gif" is available in "gif"(cid:10) format from: http://arxiv.org/ps/astro-ph/0009208v3 This figure "fig2.gif" is available in "gif"(cid:10) format from: http://arxiv.org/ps/astro-ph/0009208v3
0706.3391
1
0706
2007-06-22T19:06:23
The Overdensity in Virgo, Sagittarius Debris, and the Asymmetric Spheroid
[ "astro-ph" ]
We investigate the relationship between several previously identified Galactic halo stellar structures in the direction of Virgo using imaging and spectroscopic observations of F turnoff stars and blue horizontal branch stars from the Sloan Digital Sky Survey (SDSS) and the Sloan Extension for Galactic Understanding and Exploration (SEGUE). We show that the Sagittarius dwarf leading tidal tail does not pass through the solar neighborhood; it misses the Sun by more than 15 kpc, passing through the Galactic plane outside the Solar Circle. It also is not spatially coincident with the large stellar overdensity S297+63-20.5 in the Virgo constellation. S297+63-20.5 has a distinct turnoff color and kinematics. Faint (g ~ 20.3) turnoff stars in S297+63-20.5 have line-of-sight, Galactic standard of rest velocities V(GSR)= 130 +/- 10 km/s, opposite in sign to infalling Sgr tail stars. The path of the Sgr leading tidal tail is also inconsistent with the positions of some of the nearer stars with which it has been associated, and whose velocities have favored models with prolate Milky Way potentials. We additionally show that the number densities of brighter (g ~ 19.8) F turnoff stars are not symmetric about the Galactic center, and that this discrepancy is not primarily due to the S297+63-20.5 moving group. Either the spheroid is asymmetric about the Galactic center, or there are additional substructures that conspire to be on the same side of the Galaxy as S297+63-20.5. The S297+63-20.5 overdensity in Virgo is likely associated with two other previously identified Virgo substructures: the Virgo Stellar Stream (VSS) and the Virgo Overdensity (VOD). However, the velocity difference between the VSS and S297+63-20.5 and the difference in distance estimates between the VOD and S297+63-20.5 must be reconciled.
astro-ph
astro-ph
The Overdensity in Virgo, Sagittarius Debris, and the Asymmetric Spheroid Heidi Jo Newberg1, Brian Yanny2, Nate Cole1, Timothy C. Beers3, Paola Re Fiorentin4, Donald P. Schneider5, Ron Wilhelm6 ABSTRACT We investigate the relationship between several previously identified Galactic halo stellar structures in the direction of Virgo using imaging and spectroscopic observations of F turnoff stars and blue horizontal branch stars from the Sloan Digital Sky Survey (SDSS) and the Sloan Extension for Galactic Understanding and Exploration (SEGUE). We show that the Sagittarius dwarf leading tidal tail does not pass through the solar neighborhood; it misses the Sun by more than 15 kpc, passing through the Galactic plane outside the Solar Circle. It also is not spatially coincident with the large stellar overdensity S297+63-20.5 in the Virgo constellation. S297+63-20.5 has a distinct turnoff color and kinematics. Faint (g0 ∼ 20.3) turnoff stars in S297+63-20.5 have line-of-sight, Galactic standard of rest velocities Vgsr = 130 ± 10 km s−1, opposite in sign to infalling Sgr tail stars. The path of the Sgr leading tidal tail is also inconsistent with the positions of some of the nearer stars with which it has been associated, and whose velocities have favored models with prolate Milky Way potentials. We additionally show that the number densities of brighter (g0 ∼ 19.8) F turnoff stars are not symmetric about the Galactic center, and that this discrepancy is not primarily due to the S297+63-20.5 moving group. Either the spheroid is asymmetric about the Galactic center, or there are additional substructures that conspire to be on 1Dept. of Physics, Applied Physics and Astronomy, Rensselaer Polytechnic Institute Troy, NY 12180; [email protected] 2Fermi National Accelerator Laboratory, P.O. Box 500, Batavia, IL 60510; [email protected] 3Department of Physics and Astronomy, Center for the Study of Cosmic Evolution, and Joint Institute for Nuclear Astrophysics, Michigan State University, East Lansing, MI 48824 4Max-Planck-Institute fur Astronomy, Konigstuhl 17, D-69117 Heidelberg, Germany 5Department of Astronomy and Astrophysics, The Pennsylvania State University, University Park, PA 16802 6Department of Physics, Texas Tech University, Lubbock, TX -- 2 -- the same side of the Galaxy as S297+63-20.5. The S297+63-20.5 overdensity in Virgo is likely associated with two other previously identified Virgo substructures: the Virgo Stellar Stream (VSS) and the Virgo Overdensity (VOD). However, the velocity difference between the VSS and S297+63-20.5 and the difference in distance estimates between the VOD and S297+63-20.5 must be reconciled. Subject headings: Galaxy: structure -- Galaxy: halo -- galaxies: Sagittarius individual -- 1. Introduction A number of studies have noted that there are more spheroid stars in the Virgo constel- lation than were expected from smooth spatial models of the spheroid star density. The first indication of spheroid substructure in this region of the sky came from the Vivas et al. (2001) discovery of five excess RR Lyrae stars from Quasar Equatorial Survey Team (QUEST) sur- vey data. Although a small number, it was a significant overdensity in a small area. The QUEST collaboration later identified an excess of 23 RR Lyrae variables with this clump. These RR Lyrae stars have apparent magnitudes 16.5 < V0 < 17.5, and are within the Galac- tic coordinates 279◦ < l < 317◦, 60◦ < b < 63◦. The collaboration later dubbed this feature the "12h.4 Clump" (Zinn et al. 2004). Subsequent spectra of the RR Lyraes and BHB stars in this clump (Duffau et al. 2006) revealed a moving group at (RA, DEC) = (186◦, −1◦) with a line-of-sight, Galactic standard of rest velocity of Vgsr = 99.8 km s−1 and a distance of 19 kpc from the Sun. They suggest that this moving group be called the Virgo Stellar Stream (VSS). The existence of an overdensity of stars in the stellar spheroid near (l, b) = (297◦, 63◦) was independently confirmed by Newberg et al. (2002) using measurements of blue turnoff stars from the Sloan Digital Sky Survey (SDSS; York et al. 2000), and identified as S297+63- 20.5.1 The estimated distance of this structure from the Sun, calculated from g0 = 20.5 turnoff stars, is rSun = 18 kpc, but the color-magnitude diagram of this region does not have a tight main sequence, indicating that the structure is likely to be dispersed in distance. This overdensity appears to match the position of the "12h.4 Clump", but is somewhat offset from the VSS, which is at (l, b) = (288◦, 62◦). 1In Newberg et al. (2002) this structure was incorrectly labeled 297+63-20. The "S" signifies stream (or substructure), followed by the Galactic longitude, then Galactic latitude, and the apparent magnitude of the reddening corrected g0 turnoff. In the present paper we correct the name to be S297+63-20.5; both names will refer to the same overdensity of stars. -- 3 -- Because the mapping between the physical structure on the sky and its nomenclature in the literature for the apparent excess in the number of spheroid stars in the Virgo constel- lation is unclear, we will refrain for now from identifying the S297+63-20.5 stars with the VSS, though they are at about the same distance and position in the sky. We will return to the question of which named spheroid structures are related to each other in the discussion section of this paper. Majewski et al. (2003) also identify a structure in Virgo using the stellar catalogs from the 2MASS survey, and connect it to "M giants tens of kiloparsecs above the Galactic plane that are in the heart of the descending, foreshortened northern loop," which is their terminology for the leading tidal tail of the Sgr dwarf galaxy. Any overdensity in this part of the sky is close to the orbital plane of the Sgr dwarf galaxy, and might be related to the Sgr dwarf tidal debris. Later papers, for example Law, Johnston, & Majewski (2005), show both the leading and trailing tidal tails of the Sgr dwarf looping around through a similar position on the sky, near the identified overdensities in Virgo, before passing through the Galactic plane very close to the Sun. Since the Sagittarius tidal tails are quite prominent in the spheroid, and the observed overdensity or overdensities in the Virgo region are near the orbital plane of the Sagittarius dwarf, a connection between these structures is worthy of exploration. Juri´c et al. (2006) identify a large overdensity (the Virgo Overdensity, VOD) in the Virgo constellation, which may or may not be related to the previously named substructures. Using photometric parallaxes for a large number of SDSS stars, they show that this large overdensity is ≈ 5 − 15 kpc above the plane and covers 1000 sq. deg. of sky. They detect no downturn in the star counts toward lower Galactic latitudes, indicating that the structure could continue further into the Galactic plane than the SDSS observations probe. At that time the SDSS had not yet probed Galactic latitudes lower than about 60◦ near l = 300◦. Juri´c et al. (2006) interpret the overdensity as a significant spheroid substructure that might be an invading dwarf galaxy. The estimated distance to the VOD is smaller than the distance to the VSS or S297+63-20.5. It is unclear whether the nearer distance is real, or if distance uncertainties are large enough that the VOD, VSS, and S297+63-20.5 could describe the same physical structure. In §2 and §3 of this paper we will discuss S297+63-20.5 only; the relationship of this overdensity to the VOD and VSS will be deferred to the discussion. More recently, Martinez-Delgado et al. (2006) suggest that the invading dwarf galaxy might be the Sgr dwarf galaxy leading tidal tail, and show that the Law, Johnston, & Majewski (2005) model passes through the observed location of the VOD, and that the stellar density of the VOD is similar to the model predictions. They suggest that if the Sgr dwarf tidal stream passes through the solar neighborhood, it will be detected as an excess of stars with -- 4 -- negative radial velocities in the northern Galactic hemisphere. This paper treats the VOD as a separate entity from the VSS, but located in a similar position in space. The VOD is described by these authors as likely to be associated with the Sgr leading arm, and there- fore infalling, and the VSS is a group of stars nearly coincident in sky position, but with a different vertical and kinematic structure. Because the Sgr dwarf galaxy tidal tails can help constrain the shape of gravitational potential of the Milky Way and its dark matter halo, they are the topic of considerable study, independent from their possible connection with the observed overdensity in Virgo. One can in principle, by building ever more sophisticated equipment, know the position and velocity of every star in the Milky Way. Stars in tidal streams are the only stars which one can even in principle know about their positions and velocities in the past. Stars in tidal streams all resided, at one time in the past, in the progenitor satellite dwarf galaxy or star cluster. It is this property of tidal streams that allows us to use them to constrain the Milky Way potential. Because the Sgr dwarf tidal tails are the most prominent tidal debris in the spheroid, and because the Sgr dwarf orbital plane is approximately perpendicular to the Galactic plane, observations of the Sgr tidal debris potentially provide a strong indication of the shape of the gravitational potential surrounding the Milky Way. Interestingly, a comparison of disruption models with stellar positions and velocities in the Sgr tidal tails has produced a contradictory result. The leading and trailing Sgr tidal tails do not lie on the same plane (Newberg et al. 2003); they are tilted in opposite directions with respect to the orbital plane of the Sgr dwarf galaxy. Their tilt suggests that the Galactic potential is oblate (Martinez-Delgado et al. 2004; Johnston, Law & Majewski 2005; Law, Johnston, & Majewski 2005). However, the radial velocities of what appear to be some leading tidal tail stars can only be fit to disruption models with prolate Galactic potentials (Helmi 2004; Law, Johnston, & Majewski 2005). The leading tidal tail stars that are not compatible with oblate Galactic potentials are near the various identified substructures in Virgo. One must consider the possibility that the Milky Way potential is more complex than the models, or that some Sgr leading tidal tail stars might be misidentified. Monaco et al. (2006) seems to suggest the latter. Figure 3 of their paper shows radial velocities for leading tidal tail debris that has Vgsr measurements like those of an oblate dark matter halo, and stars with Vgsr ≈ −75 km s−1 that they interpret as arising from an ancient episode of tidal stripping. Law, Johnston, & Majewski (2005), however, show the leading tidal stream from Sgr transitioning seamlessly into a clump of stars with Vgsr ≈ −100 km s−1. It is this latter clump of Sgr leading tidal tail stars that can only be fit with a prolate halo model. The data in both papers come from 2MASS selected M giants, and the velocity dispersions of the -- 5 -- stars in the vicinity of the Virgo constellation, with Vgsr ≈ −100 km s−1, are much larger than other places along the tidal debris stream. Another complication that potentially blurs our understanding of the nature and extent of the overdensity (or overdensities) in Virgo is the suggestion that the smooth population of stars in the spheroid might be triaxial (Newberg & Yanny 2005, 2006; Savage et al. 2006). A triaxial spheroid can produce a stellar density that is not symmetric around the center of the Galaxy, as viewed from our vantage point near the Sun. This idea was proposed because several authors see more spheroid stars in quadrant IV (270◦ < l < 360◦) than we see in quadrant I (0◦ < l < 90◦), and because the lowest density of spheroid stars appears to be at a Galactic longitude slightly less than 180◦ (Newberg & Yanny 2006). A triaxial spheroid might explain how an overdensity could cover a large region of the sky and not immediately disperse; the overdensity in Virgo might be partially composed of an increased density of spheroid stars from the smooth component. Figure 3 of Newberg & Yanny (2006) shows that no model has been found that completely accounts for the observed overdensity of stars near S297+63-20.5, so although a triaxial spheroid might fit the tails of an overdensity in Virgo, we would still expect to find a more spatially concentrated component of tidal debris with coherent velocities. A key distinction between explaining an asymmetry with a 'local' stream vs. a 'global' structure such as a triaxial spheroid is the difference in observed velocity dispersion of mem- ber stars. Stars in a stream must have a relatively small velocity dispersion (σ < 30 km s−1), given the origin of the stream and given the fact that stream remains coherent as it moves through the Galaxy's potential. On the other hand, a triaxial spheroid structure in the halo may have a very large velocity dispersion (σ ∼ 100 km s−1), but evidence for its member stars will be spread around the whole of the halo. In this paper we attempt to disentangle the components of the spheroid substructure in Virgo using SDSS photometry of Blue Horizontal Branch (BHB) and F turnoff stars to build density maps of the tidal streams and the stellar spheroid itself. We show that the Sgr dwarf leading tidal tail, which is in the North Galactic Cap, arcs over (farther from the Galactic plane) the S297+63-20.5 overdensity, and over the position of the Sun in the Galactic plane. If one extrapolates the path of the Sgr leading tidal tail below b = 30◦, it passes through the Galactic plane 15 kpc or more outside the Solar Circle. This calls into question both the identification of S297+63-20.5 stars with the Sgr leading tidal tail and the identity of the M giant stars identified with the leading tidal tail of Sagittarius that have velocities in conflict with oblate models of the Galactic potential. We show that there is a clear peak in the density of 20 < g0 < 21 F turnoff stars near (l, b) = (300◦, 60◦) that is at least 20◦ across, but the total extent on the sky and distance -- 6 -- range cannot be constrained by photometry alone. The densities of F turnoff stars over a four magnitude range (18.5 < g0 < 22.5) are not consistent with an axisymmetric spheroid. If the absolute magnitude range of the F turnoff stars is 1.5 magnitudes, the spread in distance is a factor of 3, or in this case 10-32 kpc. In every direction that we probe at these implied distances, there are more stars in quadrants III & IV than there are in the symmetric position (with the same Galactic latitude but with longitude lsymmetric = 360◦ − l) in quadrants II & I. If the spheroid is axisymmetric, then the number counts should differ only where there is a significant tidal substructure. Either there is an enormous substructure centered in Virgo but covering a large fraction of the Galaxy or the smooth portion of the spheroid is not axisymmetric. We then measure the velocities of the F turnoff stars associated with S297+63-20.5 and find that their line-of-sight, Galactic center of rest velocity is Vgsr = 130 ± 10 km s−1. Throughout this paper, the symbol Vgsr will refer to the line-of-sight component of the velocity, as determined from radial velocity, transformed to the Galactic standard of rest, using a Solar motion of (vX, vY , vZ) = (10, 225, 7) km s−1 (Dehnen & Binney 1998). The measured Vgsr is not consistent with the leading tidal tail of Sgr, which has negative Vgsr, in support of our claim that S297+63-20.5 is not spatially coincident with the Sgr stream. This velocity is significantly higher than the measured Vgsr for the VSS, but not so different that the two structures could not be related in some way. In the direction (l, b) = (300◦, 55◦), we see evidence for peaks with Vgsr = −76±10kms−1 and Vgsr = −168 ± 10 km s−1. The first of the two peaks appears to have a bluer turnoff than the latter peak, and could potentially be related to the 2MASS M giant stars that have previously been assigned to the Sgr leading tidal tail. 2. Spheroid Substructure from SDSS Photometry Yanny et al. (2000), Newberg et al. (2002), and Newberg et al. (2003) show that both the leading and trailing tails of the Sgr dwarf spheroidal galaxy can be traced in SDSS color- selected BHB and F turnoff stars. Since this work, much more SDSS data have become available, covering nearly the entire North Galactic Cap. We will use the BHB and F turnoff stars in these data to trace the Sagittarius dwarf galaxy tidal tails, and explore their connection with the S297+63-20.5 overdensity in Virgo. The data presented in this paper come from the SDSS DR5 (York et al. 2000; Adelman-McCarthy et al. 2007) plus about 200 square degrees of imaging data from SDSS II that fills in the gap in DR5 photometry near the Galactic pole (to be released as part of SDSS-II DR6). We will be analyzing ugriz photometry (Fukugita et al. 1996; Smith et al. 2002) of 8500 sq. degrees in the North -- 7 -- Galactic Cap. Details of the SDSS survey geometry may be found in Stoughton et al. (2001) and Abazajian et al. (2003). Other SDSS technical information can be found in: Gunn et al. (1998), Hogg et al. (2001), Pier et al. (2002), Ivezi´c et al. (2004), Gunn et al. (2006), and Tucker et al. (2006). 2.1. BHB stars - Tracing the Sgr Tidal Tails Blue Horizontal Branch (BHB) stars are important for tracing spheroid structure be- cause they are intrinsically bright and because they are all approximately the same absolute magnitude so their distances can be estimated from apparent magnitudes to create a three dimensional picture of the spheroid. BHB stars are typically found in older stellar popula- tions such as are found in the spheroid, but are not found in all old populations. Even in the spheroid, there are a significant number of high surface gravity stars with colors that suggest a spectral type of A. These stars are either blue stragglers or young stars that have been stripped from a star-forming region such as a dwarf galaxy, and are approximately two magnitudes fainter than BHB stars. From SDSS photometry we can select stars with colors consistent with spectral type A, and then we further divide this sample into the stars whose photometry is more consistent with high surface gravity and the stars whose photometry is more consistent with a BHB star. The photometric separation of A stars by surface gravity is not perfect, but is good enough to reveal important spheroid substructures. Photometrically selected A-colored stars were chosen from the SDSS database in a similar manner to Yanny et al. (2000). The stars were selected by the criteria: "PRI- MARY," which eliminates duplicates, classified as "STAR" because the point-spread func- tion was consistent with a point source, and having colors −0.3 < (g − r)0 < 0.0, and 0.8 < (u − g)0 < 1.5. The subscript '0' indicates that the magnitudes were corrected by the Schlegel, Finkbeiner, & Davis (1998) reddening as implemented for SDSS Data Release 5 (DR5; Adelman et al. 2007). This correction is appropriate given the large distance of the BHB stars. We further separate higher surface gravity blue straggler (BS) stars from the lower surface gravity blue horizontal branch stars using the color selection criteria of Figure 10 in Yanny et al. (2000). Fig. 1 shows the spatial positions of the BHB stars in the plane of the Sgr dwarf orbit, under the assumption that all of the stars have absolute magnitude Mg0 = 0.7 (Yanny et al. 2000). We selected only BHB stars whose projected position was within 15 kpc of the Sgr dwarf orbital plane, as defined in Majewski et al. (2003), and rejected all stars within 0.2◦ of the globular clusters M53, NGC 5053, NGC 4147, and NGC 5466. The XSGR,GC and YSGR,GC coordinates are in the Sgr orbital plane, and the Galactic plane is at YSGR,GC = 0. -- 8 -- One can faintly see the newly discovered Bootes dwarf galaxy (Belokurov et al. 2006b) near (XSGR,GC, YSGR,GC) = (10, −55) kpc. This figure can be compared with Figure 11 of Majewski et al. (2003), which is reproduced in Fig. 4 of this paper, and shows the positions of 2MASS selected M giant stars in the Sgr tidal tails. In Fig. 1, one can clearly see the leading tidal tail of the Sgr dwarf tidal stream, 30-40 kpc above the Galactic plane. A piece of the trailing tidal tail is in the upper left corner. The center of the leading tidal tail appears to arc over the solar position, and to descend outside the solar circle if one extrapolates the trajectory down toward the plane. The stars in Sgr stream directly over the Sun in Fig. 1 appear to be in a very broad region of the sky, which is not consistent with results of models of Sgr tidal debris, for example those of Law, Johnston, & Majewski (2005). Since we have not surveyed the entire volume within 15 kpc of the Sgr dwarf orbital plane, there are non-trivial selection effects near the edges of the data, which in particular make it difficult to study the structure within 20 kpc of the Sun in this figure. The leading tidal tail appears unexpectedly broad in this projection as it arcs over the Sun, and the extrapolation of the leading tidal tail toward the Galactic plane is somewhat ambiguous. Since viewing the data in the observed (magnitude) rather the derived (distance) space often yields higher contrast on any structures present in the halo, we show in Figure 2 the g0 magnitude vs. Λ⊙ angular position along the Sgr tidal stream, separately for BHB and BS stars within 15 kpc of the Sgr dwarf orbital plane. The left panel contains the same stars as Fig. 1, but in apparent magnitude versus angle along the Sgr orbit. The angle Λ⊙ is measured from the center of the Sgr dwarf galaxy, as viewed from the Sun, back along the direction of the trailing tidal tail, as defined by Majewski et al. (2003). This is the longitude part of a (Λ⊙, B⊙) system in which the equator is rotated to line up with the Sgr dwarf orbital plane. A comparison of the left and right panels shows that the BHB/BS separation is quite good but not perfect. Very few of the BHB stars in the leading tidal tail of Sgr are present in the right panel showing high surface gravity (blue straggler) stars. We have not analyzed the concentration of stars brighter than g0 = 15.5. These stars are saturated in the SDSS images, and more care is required to determine how well color separation works with the interpolated magnitudes of these stars. This figure can be compared with the 2MASS M giant stars in the Sgr tidal tails of Figure 8 in Majewski et al. (2003) and Figure 2 of Newberg et al. (2003). The Λ⊙ axis is flipped from previous papers, but matches the orientation of the stream in Figure 1. There is a density peak in the left panel of Figure 2, centered at (Λ⊙, g0) = (240◦, 16.7), which we initially thought was related to at least one of the previously discovered overden- -- 9 -- sities in Virgo. However, Figure 3 shows that this overdensity is on the wrong side of the Galaxy for it to be in Virgo. Figure 3 splits the data from the left panel of Figure 2 into two pieces: the left panel of Figure 3 shows stars with ZSGR,GC < 0 and to the right shows that with ZSGR,GC > 0. Since ZSGR,GC is in approximately the same direction as the solar motion (Galactic Y ), the left panel shows primarily stars in quadrants III and IV, and the right panel shows primarily stars in quadrants I and II. The overdensity at (Λ⊙, g0) = (240◦, 16.7) is in the right panel in Fig. 3, which contains primarily stars with 0◦ < l < 180◦. The origin of this overdensity is still uncertain (but see Belokurov et al. 2007b). It could be a single debris structure or a superposition of separate streams. It is notable that several known globular clusters conspire to be in approximately the same position in this diagram; possibly there is debris associated with these globular clusters that is as yet unidentified. Note from Figures 2 and 3 that the trailing tidal tail shifts from positive ZSGR,GC to negative ZSGR,GC with increasing Λ⊙. In the left panel of Figure 3, one notes a low contrast faint extension of this tidal tail down to (Λ⊙, g0) = (250◦, 17.3). In the right panel of Figure 3 one sees the leading tidal tail at positive ZSGR,GC. This also decreases substantially in density as one proceeds along it to the left, and appears to be close to ZSGR,GC = 0 so it is split between the left and right panels in Figure 3. In both the leading and trailing tidal tails, we see the pile-up of debris near apogalacticon. In this projection, since the Sagittarius dwarf galaxy is 14◦ below the Galactic plane, material falling down on us from close to the North Galactic Pole in this orbital plane would fall toward us at Λ⊙ = 270◦ − 14◦ = 256◦. Since the leading tidal tail arcs past this angle, it likely passes through the Galactic plane outside the solar circle. 2.2. F turnoff Stars - Separating Sgr from S297+63-20.5 Since there are many more F turnoff stars in any stellar population than there are BHB stars, these F stars are in principle better tracers of substructure. The disadvantages of using F turnoff stars is that they are 3 or 4 magnitudes fainter than the horizontal branch, and have a broader spread in absolute magnitude (so they are less accurate distance indicators). The Sagittarius dwarf galaxy and tidal stream have a particularly blue turnoff which makes them easy to separate from other Galactic populations. In this section we present data which follows the leading tidal tail from the Sagittarius dwarf galaxy and shows that it misses the position of the S297+63-20.5 overdensity and also passes over the position of the Sun before falling down on the Galactic plane outside the solar circle. We then present polar plots of F star density in the North Galactic Cap in the magnitude ranges 19 < g0 < 20, 20 < g0 < 21, and 21 < g0 < 22. The polar plots show that the Sagittarius leading tidal tail is fainter -- 10 -- and offset from the S297+63-20.5 overdensity, and also show the extent of the S297+63-20.5 overdensity. Fig. 4 shows our detections of Sagittarius dwarf tidal debris from F turnoff stars in SDSS stripes 13 and 15-23, along with the previous detections of Sgr debris from BHB stars (Figure 3 of Newberg et al. 2003) and 2MASS M giants (Majewski et al. 2003). The new Sgr debris positions were determined by selecting by eye the center of the highest density peak in a histogram of g0 versus angle along the SDSS stripe for blue F stars (0.0 < (g − r)0 < 0.3). These positions were then converted to (XSGR,GC, YSGR,GC, ZSGR,GC), assuming Mg0 = 4.2. This absolute magnitude was determined in Newberg et al. (2002) by comparison with the Sgr BHB stars, so the SDSS BHB star distances should be on the same scale as the SDSS F turnoff star distances, though there could be an overall scale error. Note that the Sagittarius M giants have systematically smaller heliocentric distances than either the A stars or the F stars (Chou et al. 2006). The debris was identified as Sgr tidal debris because it was contiguous with the over- density in the adjacent stripe. There was some freedom in choosing the stream positions, as the profile of the overdensity was often asymmetric or bifurcated (Belokurov et al. 2006a). The larger open squares in Figure 4 show the positions of the higher density tidal debris. If there was a second piece of tidal debris near the main Sgr leading tidal tail in any stripe, it is shown in Figure 4 as a smaller open square. The positions of the Sgr F turnoff star detections are presented in Table 1. Fig. 4 clearly demonstrates that the leading tidal tail of the Sagittarius dwarf will pierce the Galactic plane outside the solar circle. This is in contrast to the 2MASS M star observations shown on the same figure. It is known that the stellar population in the tidal tails is more metal poor and older than the stars in the core of the Sgr dwarf galaxy (Bellazini et al. 2006), so it is reasonable to expect that the stellar populations would differ as a function of distance along the stream. However, one doesn't expect that stellar population differences would produce this much of a distance difference between estimates from F turnoff stars and M giants. Another possibility that should be considered is that the M giants near the end of the leading tidal tail are not all related to Sgr. Fig. 4 also shows that although S297+63-20.5, indicated with a large cross, is close to the plane of the Sagittarius dwarf orbit, S297+63-20.5 is much closer to the Sun than the Sgr leading tidal debris. Any debris that is in the Sgr dwarf orbit is plausibly connected to this dwarf galaxy, but this significant overdensity in Virgo is not an extension of the Sgr leading tidal tail, and there is no obvious connection with Sgr. The distance discrepancy between S297+63-20.5 and the Sgr leading tidal tail is demon- -- 11 -- strated in another way by looking at polar plots of F star densities in the North Galactic Cap at different apparent magnitudes. We selected the stellar data from SDSS DR6 with 0.2 < (g − r)0 < 0.3 and (u − g)0 > 0.4. The (g − r)0 range was selected to be centered on the turnoff color of S297+63-20.5, which is measured as 0.26 in Newberg et al. (2002). By excluding stars stars redder than (g − r)0 = 0.3, one avoids thick disk turnoff stars, even if they exist at these faint magnitudes. Fig. 5 (lower panel) shows F stars with 20 < g0 < 21; this magnitude range brackets the estimated apparent magnitude of the turnoff of S297+63- 20.5. The figure includes photometric data from SDSS, as well as some extra photometric data obtained as part of the Sloan Extension for Galactic Understanding and Evolution (SEGUE), which consists of 2.5◦-wide scans at constant Galactic longitude which appear as radial extensions to our sky coverage of the North Galactic Cap in this figure. The stars in Fig. 5 trace structure in the spheroid of the Milky Way at distances ∼ 14.5 − 23 kpc from the Sun. A larger density of F stars is represented by a darker shading in the equal-area polar plot, centered on the North Galactic cap. The Sagittarius leading arm is clearly visible running from (l, b) = (205◦, 25◦) to (305◦, 65◦). More prominent is the S297+63-20.5 overdensity, centered near l = 300◦, and hugging the lower edge of the data. Note that although there could be some overlap in the spatial positions of some of the stars, the denser portions of the Sgr stream are separate from S297+63-20.5. While most of the stars in the S297+63-20.5 are contained within an area of the sky 15◦ in diameter, the tails of the distribution could cover half of the North Galactic Cap. In order to show conclusively that the polar plot cuts through S297+63-20.5, we perform the mathematical equivalent of folding the data from the top half of the lower panel in Fig. 5 down over the line described by l = 0◦, l = 180◦, and subtracting it from the data on In other words, data between l = 0◦ and l = 180◦ is subtracted from the bottom half. the symmetrical point with the identical b at l′ = 360◦ − l and the residual is binned and plotted. The result, displayed in the top panel in Fig. 5, shows a localized structure near (l, b) = (300◦, 64◦), which is at the same position as S297+63-20.5, within the errors of each measurement. This folding of the data should leave zero residual (except where there are streams or other halo substructure) if the spheroid is axisymmetric. If the spheroid is asymmetric, then more stars should have been subtracted over a wide area of the upper panel, but it is difficult to make a smooth spheroid density function that would include a strong peak like the one we see at l = 300◦. To clarify the relationship between S297+63-20.5 and Sgr, we show in Fig. 6 the F star polar plots for stars selected with 19 < g0 < 20 and 21 < g0 < 22. The upper panel shows the brighter stars. There is still a clear excess of stars at l > 180◦ compared to the position in the sky that is symmetric with respect to the Galactic center. For example, there are -- 12 -- many more stars at (l, b) = (285◦, 60◦) than there are at (l, b) = (75◦, 60◦). However, there is not a clear peak in the star counts. The asymmetry in star counts could either be an asymmetry in a smooth distribution of stars or a dispersed stream over a large area of sky. The lower panel is dominated by a more distant piece of the Sagittarius dwarf leading tidal tail, crossed at b ∼ 50◦ with the much fainter Orphan Stream (Belokurov et al. 2007a). There are excess stars near (l, b) = (300◦, 60◦), but they are confused with and overwhelmed by stars from the Sgr tidal tails at this magnitude. A connection between the Sgr tidal tail and S297+63-20.5 cannot be completely ruled out, but the bulk of the turnoff stars in the Sgr stream are at least a magnitude fainter (g0 ∼ 21.5) than the bulk of the turnoff of S297+63-20.5 (g0 ∼ 20.5). Models of Sgr debris such as that of Law, Johnston, & Majewski (2005) show some debris stars near the position of S297+63-20.5, but not a strong peak of debris. Either the models do not adequately match the stellar debris, or S297+63-20.5 was not stripped from Sgr. The polar plots show in a compelling way that the Sgr leading tidal tail intersects neither the Sun nor the S297+63-20.5 overdensity in Virgo. The turnoff stars in the Sgr dwarf tidal stream are fainter than g0 = 21.0 near (l, b) = (297◦, 63◦) - they show up only in the faintest polar plot on the lower panel of Fig. 6. The Sgr leading tidal tail is also separated from S297+63-20.5 by about ten degrees in the sky (with significant overlap). As the leading tidal tail descends closer to the Sun we see the stream moving toward the anticenter (Fig. 5). In other words, it passes over the Sun in the North Galactic Cap. Close to the anticenter, the Sgr stream is fainter than g0 = 21 (compare Fig. 5 with Fig. 6 near the anticenter). Near (l, b) = (205◦, 30◦) the Sgr stream is visible in all of the F turnoff star polar plots, but the star density is the greatest at this position for 20 < (g − r)0 < 21. Since the the Str stream turnoff at l = 205◦ is at g0 = 20.5, its implied distance is 18 kpc from the Sun. The Sgr stream is therefore at least 15 kpc from the Solar position and well beyond the Solar Circle as it descends toward the Galactic plane. 2.3. F Turnoff Stars - separating S297+63-20.5 from the Smooth Spheroid We have already remarked that S297+63-20.5 seems to extend over a large range of magnitudes and a significant fraction of the sky. We now study whether that large extent could be partially explained by an asymmetric (triaxial) smooth distribution of spheroid stars. Fig. 7 shows quantitative comparisons between the numbers of F turnoff stars as a function of magnitude near S297+63-20.5 and those at a symmetric position in the Milky -- 13 -- Way, with the same Galactic latitude but on the opposite side of the Galactic center. If the spheroid is axisymmetric, then one could find the number of stars in S297+63-20.5 from the difference between the star counts at (l, b) = (297◦, 63◦) and the star counts at (l, b) = (63◦, 63◦). In particular, Fig. 7 shows a comparison of star counts at two places in Virgo: (l, b) = (288◦, 62◦) (upper panel) and (l, b) = (300◦, 55◦) (lower panel), with the corresponding position on the sky on the other side of the Galactic center. These two positions were selected because they match the positions of SEGUE spec- troscopic fields in which we will analyze radial velocity kinematics later in this paper. Each sky area probed has a radius of 1.5◦, which matches the footprint of a SEGUE spectroscopic plate. In addition, F stars are selected with a color range: 0.2 < (g − r)0 < 0.4 to match the color range with the highest concentration of F stars in S297+63-20.5. This redder color range will include a significant number of thick disk stars at magnitudes brighter than g0 = 18, but at fainter magnitudes there will be little contamination. Since the thick disk (unlike the halo), is axisymmetric, there should be no effect on the difference in bright star counts due to thick disk stars between the Virgo selected regions and the symmetric reference regions. Comparing the star counts at (l, b) = (288◦, 62◦), which is centered on the central knot of the VSS, with the star counts at (l, b) = (72◦, 62◦) in the upper panel of Fig. 7, one sees that the star counts start to diverge at about 18th magnitude in g0. By g0 = 20.5, there are not quite a factor of two more stars in the VSS field than in the comparison field. In the magnitude range 19.4 < g0 < 20.0, there are 560 stars in the (l, b) = (288◦, 62◦) field and only 357 stars in the corresponding reference field at (l, b) = (72◦, 62◦). Of the stars in the VSS field, 36±6% are part of S297+63-20.5. In the magnitude range 20.0 < g0 < 20.3, there are 369 stars in the VSS field and 206 stars in the reference field. Of the fainter stars in the VSS field, 44±7% are part of S297+63-20.5. This means that if one selected a random turnoff star in the VSS field at g0 = 20.5, there is nearly a 50% chance of it being a star from S297+63-20.5. The (l, b) = (300◦, 55◦) field in Virgo diverges from the reference field at (l, b) = (60◦, 55◦) at the even brighter magnitude of g0 = 17 (lower panel of Fig. 7). About a quarter of the stars with 17 < g0 < 19 are part of the S297+63-20.5 excess in Virgo. In the magnitude range 19.4 < g0 < 20.0, there are 699 stars in the Virgo field, and only 440 in the reference field. Of the stars in the Virgo field, 37±5% are part of S297+63-20.5. In the magnitude range 20.0 < g0 < 20.3, there are 453 stars in the Virgo field and 238 stars in the reference field. Of the fainter stars in the Virgo field, 47±6% are part of S297+63-20.5. At a magnitude of g0 = 20.5, more than half of the stars in the (300◦, 55◦) field should be part of the excess. If the excess population has a unique kinematic signature, as is expected for a stream or -- 14 -- satellite, it should be quite prominent in any sample of faint F turnoff stars. We will return to these fractions when we examine the kinematic distribution of selected F turnoff star spectra in these fields. The wide range of magnitudes and sky coverage makes the overdensity in Virgo difficult to rationalize with our expectations for tidal debris, even before we begin looking at the spec- troscopy results. We consider a non-axisymmetric spheroid model as a possible alternative explanation for the excess star counts in the Virgo region. To compare the observed star counts with those predicted by the best currently avail- able asymmetric model of the spheroid, the Galactocentric triaxial Hernquist model of (Newberg & Yanny 2005, 2006; Savage et al. 2006) is chosen. In this model, the center of the spheroid distribution is fixed at 8.5 kpc from the Sun with an assumed absolute mag- nitude of blue F turnoff stars of Mg = 4.2. If the absolute magnitude is incorrect, then the overall scale of all lengths and distances in the model will be off by the same factor. See Newberg & Yanny (2006) and Savage et al. (2006) for parameter definitions and values for the triaxial Hernquist model. The model was generated to fit SDSS F turnoff star counts with color 0.1 < (g − r)0 < 0.3, which is a bit bluer than the data at 0.2 < (g − r)0 < 0.4. To account for this, the normalization of the model is adjusted upward by a factor of 1.5 so that it matches the star counts as well as possible. We cannot correct for the fact that the model does not include the thick disk stars at brighter magnitudes, so we will have to ignore the difference between the star counts and the model brighter than about g0 = 18. These triaxial spheroid model curves are plotted in Fig. 7 for comparison with the F turnoff star counts. The model is a reasonable fit to the data for 18 < g0 < 21. At brighter magnitudes the thick disk dominates, and at fainter magnitudes there are fewer stars than the model predicts. The discrepancy at faint magnitudes is probably due to a deficiency in the models, and possibly the lower completeness of the fainter stellar data. Note that if one compares the star counts at (l, b) = (288◦, 62◦) and (l, b) = (300◦, 55◦) with the model, the star counts begin to diverge somewhere between g0 = 19 and g0 = 20. There is a much smaller excess in the number of stars at magnitudes brighter than g0 = 20. Even using the current best fit triaxial spheroid model, however, one cannot account for all of the excess stars, particularly those with magnitudes 20 < g0 < 21. In the next section the two possibilities for explaining the stellar count excess are ex- plored using additional information obtained from a sample of spectra from the same pop- ulation of stars. If the spheroid is axisymmetric, and the S297+63-20.5 structure is tidal debris, then 37±4% of the stars with 19.4 < g0 < 20.0 should have coherent velocities (i.e. show a small σ < 30 km s−1 dispersion), and 46±5% of the stars with 20.0 < g0 < 20.3 -- 15 -- should have coherent velocities. If the spheriod is asymmetric, then the fraction of stars with coherent velocities is expected to be lower. With the best fit model presented, one estimates that at (l, b) = (288◦, 62◦), 64 of 560, or 11% of stars with 19.4 < g0 < 20.0 should have coherent velocities, and 159 of 369, or 43% of stars with 20.0 < g0 < 20.4 should have coherent velocities. In the direction (l, b) = (300◦, 55◦), 76 of 699, or 11% of the stars with 19.4 < g0 < 20.0 should be coherent, and 127 of 453, or 28% of stars with 20.0 < g0 < 20.4 should be coherent. The model numbers are an estimate, and will change somewhat as new non-axisymmetric spheroid models are developed. 3. Substructure in Velocities of F Turnoff Stars Since S297+63-20.5 was discovered in F turnoff stars, and is prominent against the background of smooth spheroid stars in that population, one would like to measure the velocities of these same types of stars if possible. Since the overdensity is most noticeable in quite faint F turnoff stars (with 20.0 < g0 < 21.0), this is not an easy task. Essentially zero F turnoff stars in this magnitude range were observed in the primary SDSS galaxy survey. We are currently in the middle of SDSS II, which includes three survey projects, includ- ing the Sloan Extension for Galactic Understanding and Exploration (SEGUE). The SEGUE portion of SDSS II is designed to reveal velocity substructure in the Milky Way thick disk and spheroid through analysis of radial velocities and stellar properties from spectroscopy of ∼ 250, 000 Galactic stars. SEGUE spectra are identical to SDSS spectra, in that they are observed with the same instrument and processed with the same (evolving) software; the spectra cover 3700-9300 A with a resolution of ≈ 2000. The difference is that SDSS targeted a very few Galactic stars, most of which are BHBs or bright F subdwarfs that were used as standard stars, while SEGUE targets Galactic stars exclusively. While the SDSS obtained galaxy spectra in all areas of the sky with imaging data, SEGUE samples the sky with a set of pencil beam spectral surveys in 200 target fields, spaced roughly 10 to 20 degrees apart over 3/4 of the sky. In the SEGUE survey, two 640 fiber plates of spectra (Stoughton et al. 2001) are ob- tained in each target field. The brighter targets are put on one plate which is exposed about 45 minutes. Stars fainter than about r0 = 17.8 are assigned to the second plate, which is exposed for 1.5 to 2 hours. The faint plates reach about g0 = 20.5, which is just at the bright end of the S297+63-20.5 turnoff population. Targets for SEGUE spectroscopic fibers are chosen by color from the SDSS imaging data to represent a variety of stellar types at a variety of distances sampling the thick disk and spheroid of the Milky Way. -- 16 -- We have already obtained two pairs of SEGUE spectroscopic plates in directions that probe the S297+63-20.5 overdensity. These plate pairs are numbered 2689/2707 (bright/faint) at (l, b) = (300◦, 55◦) and 2558/2568 at (l, b) = (288◦, 62◦). The second plate pair is centered on the VSS RR Lyrae star overdensity. During DR6 spectro test processing of globular cluster calibration stars with known cataloged velocities, a universal offset of 7 km s−1 was noted in the wavelength solutions; this zero point correction was added to all SEGUE radial velocities used in this paper. Most of the faint stars of interest, with 0.1 < (g−r)0 < 0.4, are selected as "F subdwarf," "Blue Horizontal Branch," or "F/G" spectroscopic target types. The BHB category goes as red as (g − r)0 < 0.2 and as faint as g0 < 20.5. These stars are randomly sampled to favor lower surface gravity stars using a photometric measure of surface gravity for A-colored stars called the v-index (Lenz et al. 1998) to estimate surface gravity. The F/G category randomly samples stars with r0 < 20.2 and 0.2 < (g − r)0 < 0.48. The faint subdwarf category favors F stars with low metalicity by favoring stars that are at the very bluest tip of the stellar locus in (u − g)0 and (g − r)0, using a measure of position along the stellar locus in color space called P 1 (Helmi et al. 2003). The criteria for selecting each SEGUE category are complex, and include techniques for varying the likelihood of selecting a star for spectroscopy as a function of apparent magnitude, so that we select fewer of the more numerous, fainter stars. The broader goal is to sample all of the components of the Milky Way, so less common components are favored in the selection process. Fig. 8 shows the color magnitude diagram of the bluer SDSS stars within 1.5◦ of the center of the SEGUE plate pair 2689/2707, along with the colors and magnitudes of the stars with 0.1 < (g − r)0 < 0.4 and 19.4 < g0 < 20.5 for which we have obtained SEGUE spectra with the two SEGUE plate pairs 2689/2707 and 2558/2568. This is the entire region from which the spectra in this paper are selected. The symbols show that the spectra are spread in color and magnitude, and do not follow the color distribution of the spheroid stars. We must select stars in narrow color and magnitude ranges that correspond to a component we intend to sample; we cannot rely on the fact that one component has many more stars to guarantee that it will dominate the sample. The stars that we expect to best sample the turnoff of S297+63-20.5 are those with 0.2 < (g − r)0 < 0.4 and 20 < g0 < 21. The SEGUE spectra only go as faint as g0 < 20.5, and most of the spectra have g0 < 20.3, so we sample only the brightest end of the magnitude range. The right panels of Fig. 9 show Vgsr for stars from SEGUE faint plates 2707 and 2568 (available publicly with the release of SDSS DR7) that have the colors and magnitudes of S297+63-20.5 F turnoff stars, and radial velocity errors, as measured by cross-correlation with ELODIE standard templates (Moulataka, et al. 2004), of less than 20 km s−1. -- 17 -- In both plates, there is a clear excess of faint stars with positive radial velocities. In the combined panel (bottom right of Fig. 9), there is a five sigma peak in the bin 112 < Vgsr < 144 km s−1. There are 30 stars with positive Vgsr and only 13 stars with negative Vgsr. If all of the excess stars are associated with S297+63-20.5, then 40±11% of the fainter F turnoff stars could be associated with a single tidal debris structure in Virgo. This fraction is consistent with estimates from photometry, whether or not the spheroid is symmetric. In the symmetric spheroid case, we expected 46 ± 5% of the excess stars with 20.0 < g0 < 20.3 to be in the excess population. If the triaxial spheroid model is correct, the percentage in the excess population is about 35% for both plates combined. Our measured value of 40 ± 11% is consistent with either of these options. Therefore, from the faint (g0 > 20) spectroscopic and photometric data combined we conclude: the majority of the stars in the S297+63-20.5 overdensity are consistent with the presence of single substructure with a coherent velocity of 130 ± 10 km s−1 and relatively small dispersion, as we would expect for tidal debris. In the left panels of Fig. 9, we show Vgsr for the stars with 19.4 < g0 < 20.0. In these plots, there is no excess in the 122 < Vgsr < 144 km s−1 bin. One plate has more positive velocities, and the other has more negative velocities. In the combined panel (lower left of Fig. 9), the only significant peak has −176 < Vgsr < −144 km s−1. Thus the brighter F stars, in the left panels of Fig. 9, have distinctly different kinematics than the fainter stars in the right panels of the same figure. Out of 72 brighter stars, a few at best are part of the moving group we identified in fainter stars. The symmetric spheroid assumption indicated that we should have found 37%, which is 27 stars, in a peak near 122 < Vgsr < 144 km s−1. The most we can imagine assigning to a peak are about seven stars from the plate at (l, b) = (288◦, 62◦) that are around Vgsr = 200 km s−1. The moving group is ten percent or less of the stars with 19.4 < g0 < 20.0. This differs at the 4σ level with our expectation of 27 stars present in a peak if the halo is axisymmetric. This measurement of brighter (19.4 < g0 < 20) star kinematics and photometry thus favors a triaxial spheroid. In the symmetric spheroid model (plus individual star stream peaks), we expected 37 ± 4% of the stars to be coherent in velocity, and we only found 11% at the most. The triaxial spheroid model predicted that 11% of the stars would be coherent in velocity, which is in good agreement with the observed fraction. There is one significant peak in the velocities of the brighter F turnoff stars, but it has a very negative Vgsr. We identify as a "peak" any bin with more than a 2.5σ excess in the expected number of stars, where σ2 = y(1 − y/n), y is the expected number of stars, and n is the total number of stars in the 100 square degree region. The limit for a bin to be identified as a peak is shown by the dotted curve in Fig.9. In plate 2707, there are five stars -- 18 -- with −176 < Vgsr < −144km s −1 where we would have expected only one, and three stars in plate 2568 where we would have expected only two. This amounts to an excess of about five stars in the summed histogram, lower left of Fig. 9. The mean velocity of this peak is −168 ± 10 km s−1. Since Newberg et al. (2002) measure a turnoff color for S297+63-20.5 of (g − r)0 = 0.26, one would not expect many stars in this structure to have measured colors of 0.1 < (g −r)0 < 0.2 in the SDSS. We show measurements of Vgsr for these bluer stars in Fig. 10. In the fainter (20.0 < (g − r)0 < 20.5) set of stars, there are maybe two excess stars in the bin one to the right of the moving group associated with the center of S297+63-20.5, but no significant velocity peaks are observed. In the left panels of Fig. 10, we see two marginally significant peaks: one at nearly the same velocity as the Vgsr = 130 km s−1 peak in Fig. 10, and one at in the bin with −80 < Vgsr < −48. We looked at the individual stars in positive velocity peak, and most of them are at the very faint and very red edges of the selection criteria. Possibly, they represent blue stragglers that are at the same distance but have brighter absolute magnitudes than the F turnoff stars in the moving group identified in the fainter panels of Fig. 9. The peak with negative radial velocity is at −76 ± 10 km s−1, but note that there are more stars on the negative velocity side of this peak than on the positive side. If most of the brighter stars in this color range are associated either with the positive or the negative Vgsr peak, then the Gaussian distribution from the expected spheroid would be much lower and the center of the peak would move one bin to the left. The center of the peak is nearly between two bins. This peak is more pronounced in plate 2707 than in plate 2568. Once this peak is identified, one can see that it also present at some level in the fainter panels on the right of Fig. 10. Since the stars in this peak are at about the same distance and line-of-sight velocity as stars which other authors have attributed to the Sgr dwarf leading tidal tail, they will be analyzed at greater length in the discussion section of this paper. 4. Discussion 4.1. Is S297+63-20.5 Related to the Sagittarius Dwarf Tidal Stream? There was a time when any significant clump of tidal debris in the spheroid was suspected of belonging to the one known example of present-day merging in our galaxy − the Sagittarius dwarf tidal stream. We now know there are many tidal debris streams, though the only currently known tidal stream as large as that of the Sagittarius dwarf is the Monoceros stream in the Galactic plane (Yanny et al. 2003; Ibata et al. 2003). We identify stars in the -- 19 -- tidal tails of the Sagittarius dwarf spheroidal galaxy because either their locations and radial velocities are contiguous to those of other known pieces of Sagittarius debris or because their locations and velocities match those of models. In this paper, we have presented a strong argument that S297+63-20.5 is not part of the Sgr leading tidal tail because the turnoff stars in S297+63-20.5 are much brighter than those in the leading tidal tail debris in the same direction in the sky, and the stars in S297+63- 20.5 have positive Vgsr while Sgr leading tidal debris would have negative Vgsr. Moreover, Newberg et al. (2002) showed that the turnoff of S297+63-20.5 is redder than that of the Sgr tidal stream anywhere it has been detected in the sky. Although Martinez-Delgado et al. (2006) were able to generate a model that shows the leading tail of the Sagittarius stream going through the position of the S297+63-20.5, they did so at the cost of not fitting some of the more distant spatial detections of that same tidal tail. In this paper, (see §2), we locate additional pieces of what appears to be the leading tidal tail of Sagittarius in SDSS F turnoff stars, and show that it does not intersect the position of S297+63-20.5. Since the Sgr tidal debris may be bifurcated, lumpy, wrap multiple times around the Milky Way, and have different stellar populations in different places, it is difficult to prove that any spheroid substructure, especially one that happens to be in the Sgr orbital plane, is unrelated to this tidal disruption event. Since the S297+63-20.5 debris is outgoing in velocity, it is possibly related to the Sgr trailing tidal tail. However, no existing model shows a lump of material near S297+63-20.5. More extensive modeling and mapping of the Sgr tidal debris is required to determine definitively whether they are related to each other. 4.2. Relationship between S297+63-20.5 and other Virgo Substructures What is much murkier is the relationship between S297+63-20.5 and the other substruc- tures that have been discovered in the Virgo region, and have been identified as overdensities in Virgo. These include the VSS (aka the "12h.4 Clump"), and the VOD. The VSS is in the same general area of the sky as S297+63-20.5, and at a similar calculated distance. They both have positive Vgsr, though the VSS has a measured Vgsr of 99.8 km s−1 compared to our measurement of 130 km s−1 for S297+63-20.5. It is possible that membership or measurement errors of a few stars in either the S297+63-20.5 sample or the VSS sample could explain the difference in measured velocity. The central knot of the VSS is also a bit offset (9◦) from the center of S297+63-20.5. One wonders whether the clump of RR Lyrae stars that defines the VSS could be a substructure within S297+63-20.5, -- 20 -- such as a disrupted GC within a disrupted dwarf galaxy. Or, we may find as the number of known stars in the VSS grows, it might resemble the S297+63-20.5 structure more closely. Given the prevalence of velocity substructures in even the small amount of data analyzed here, it cannot be completely ruled out that the VSS and S297+63-20.5 might be chance superpositions. However, that possibility seems fairly unlikely since the velocities are quite similar. We expect that as more data is analyzed, the S297+62-20.5 overdensity will become known as the Virgo Stellar Stream (VSS). It is somewhat difficult to assess the relationship between S297+63-20.5 and the VOD. In the Juri´c et al. (2006) paper, their first step is to determine the distance to each star in the SDSS photometric database. All subsequent analyses are carried out with these spatial positions. The VOD has a distance of ∼ 5 − 15 kpc, which is a little closer than our estimate of 18 kpc for S297+63-20.5, which is at a similar estimated distance as the VSS. However, our distance errors to S297+63-20.5 are large, and there is no guarantee it is the same as the VSS. In particular, there is no guarantee that our absolute magnitude estimate of Mg = 4.2, which was derived for Sgr tidal debris with similar colors, is appropriate for S297+63-20.5. The factor of 2 density excess for the VOD is similar to the density excess that we see in comparisons of the number of F turnoff stars near (l, b) = (300◦, 60◦) and (l, b) = (60◦, 60◦). Also supporting the identification of S297+63-20.5 with the VOD is the position in the sky and the large sky area covered. If there were two separate overdensities, one at 10 kpc and one at 18 kpc, we would expect to see a separate VOD structure in the top panel of Fig. 6, with an apparent magnitude around g0 = 19.2. All we see in that figure is a general asymmetry in the star counts - no detected peak density. This suggests that either our technique of estimating distance gives different results from Juri´c et al. (2006) or that the turnoff of the VOD is redder than (g −r)0 = 0.3. Probably, they are the same overdensity measured with the same star catalog, but it is difficult to know for sure. What is needed is a representation of the VOD stars that allows us to identify which stars are members of the structure, such as a color magnitude diagram and a map of angular position on the sky. Then we would be able to check whether we are looking at the same stars in the same catalog or not. 4.3. The Sgr Leading Tidal Tail and the Shape of the Milky Way Gravitational Potential The controversy regarding the oblateness of the Milky Way gravitational potential hinges on the radial velocity data for M giants in the part of the leading tidal tail that is closest to the Sun, 240◦ < Λ⊙ < 260◦. Figure 12 from Law, Johnston, & Majewski (2005) shows the -- 21 -- model distances and velocities of Sagittarius debris for prolate, spherical, and oblate dark matter halos, along with data for 2MASS M stars. There are a large number of stars near Λ⊙ = 250◦ that have velocities of −150 < Vgsr < −50 km s−1 and distances from the Sun of about 15 kpc. It is precisely these stars that do not fit the published models, unless one assumes a very prolate dark matter halo (Helmi 2004). The stars with 200◦ < Λ⊙ < 275◦ also have a larger velocity dispersion than Sgr debris stars at other places in the stream. The last column in Table 1 gives the position of the Sgr leading tidal tail in terms of Λ⊙; at Λ⊙ = 250◦ the distance to the Sgr stream is 36 kpc. M giant distances in Majewski et al. (2003) are systematically 13% less than ours, as determined from a compar- ison of M giant stars and A/F stream position of the leading tidal tail stars near apogalac- ticon (Newberg et al. 2003). Applying this scale difference, the M giant stars at Λ⊙ = 250◦ should be at 31 kpc on the M giant distance scale. There are very few stars in the Law, Johnston, & Majewski (2005) sample that are that far away. Fig. 4 shows where the discrepancy arises. If one looks directly up from the Sun in Fig. 4, following the direction of −YSGR,GC, the Sgr F turnoff stars are estimated to be more than 30 kpc above the Galactic plane. In contrast, the M giant stars are scattered from 10 kpc to 35 kpc in the same direction. The M giant stars are at about the same distance, and in about the same direction, as the S297+63-20.5 overdensity. In fact, we detect a radial velocity peak for F turnoff stars near S297+63-20.5 that has a velocity of Vgsr = −76 km s−1, which is very similar to the M giant velocities. There is an excess of stars in the bin with −80 < Vgsr < −48 km s−1 in every histogram in Fig. 10. There are 15 of these bluer F turnoff stars (0.1 < (g − r)0 < 0.2) with 19.4 < g0 < 20.5 where we expected to see less than seven. The excess in this bin is more than 3 sigma, and there are additional extra blue F turnoff stars with more negative Vgsr. There is also velocity peak in RR Lyrae stars near the VSS, at about the same distance and direction. Figure 2 of Duffau et al. (2006) shows spectroscopy of 28 BHB and RR Lyrae stars from the QUEST survey. In the Duffau et al. (2006) paper, a very tight group of stars at 18.5-20 kpc from the Sun and within 10◦ of each other in the sky is identified. The mean velocity in the Galactic rest frame of these stars is 99.8 km s−1. Additional stars as close as 16 kpc and as distant as 24 kpc have similar enough radial velocities that they might be considered to be part of the same kinematic structure. The remaining stars are not consistent with a Gaussian distribution, and we note that there is a very significant peak in the lower panel of their Figure 2 that is very near Vgsr = −75 km s−1. If one considers all of the radial velocities obtained by Duffau et al. (2006), there are 12 stars in a peak at nearly 100 km s−1, nine stars in a peak at ≈ −80 km s−1, and -- 22 -- 7 stars with very large negative or positive radial velocities that are consistent with neither peak, nor with any existing model of spheroid kinematics. There is insufficient information in the Duffau et al. (2006) paper to adequately analyze the nine stars with velocities toward the Sun, but they likely cover an area 20◦ on the sky or larger, and at least the distance range 14 < rSun < 20 kpc. These stars are very likely associated with our F turnoff star peak at Vgsr = −76 km s−1 and with the stars that Law, Johnston, & Majewski (2005) assigns to the Sgr leading tidal tail. Figure 3 of Monaco et al. (2006) also show stars in the same part of the sky, Λ⊙ = 239◦, and with similar velocities, Vgsr ∼ −100 km s−1. Interestingly, they identify these stars as part of an ancient tidal debris stream that was stripped from Sgr > 2 Gyr ago. Although this presents a possible explanation, it does not explain why the dispersion of the stars appears so large in this portion of the sky, and is at odds with the general assumption that 2MASS giants favor younger, more recently stripped portions of the debris. Figure 12 of Law, Johnston, & Majewski (2005) shows a subset of stars in this region that have a very narrow velocity dispersion and very nearly match the expected position of the ancient tidal debris if one assumes an oblate dark matter halo. One wonders whether a simpler explanation for the debris with Vgsr ≈ −100 km s−1 might be the presence of a previously unidentified spheroid component. The fact that the stars in this infalling structure have a blue turnoff supports, but does not prove, the claim that they are part of the Sgr dwarf tidal stream. However, they appear to be closer than the main part of the tidal tail. Further investigation is required to determine whether they are part of the Sgr dwarf tidal debris, and whether they represent the population of stars they are being compared to in the tidal disruption models. 4.4. The Shape of the Spheroid If we knew the spatial position of every spheroid star, or at least a representative sample of spheroid stars, we could use this information to construct a density model for this Milky Way component. But given that there are very large known overdensities in the spheroid, what would the parameters in this model be? Bell et al. (2007) show that all smooth models are poor fits to the distribution of stars in the spheroid. We have chosen to model the spheroid as a set of significant overdensities that are coherent in position and velocity, plus a roughly smooth spheroid distribution. Overdensities include globular clusters, dwarf galaxies, and large chunks of associated debris from these objects. Thus, we are now in the process of identifying large, coherent debris structures so -- 23 -- that we can fit the remaining stars to a smooth spheroid model; and fitting smooth spheroid models so we can identify overdensities that might be large, coherent debris structures. There is no controversy over the asymmetry of the star counts in the stellar spheroid stars in the North Galactic Cap (Newberg & Yanny 2005, 2006; Xu, Deng & Hu 2006), though there is a question of whether there is an asymmetry in the south (Xu, Deng & Hu 2007). An important question is whether the asymmetry is due to a single large debris structure added to an otherwise axisymmetric distribution of stars. The data in this paper suggest that is not the case. The S297+63-20.5 is a very sig- nificant, coherent overdensity that contains about half the spheroid stars at g0 = 20.5 and (l, b) = (297◦, 63◦). However, we have shown that it is localized in magnitude and velocity, and does not explain the asymmetry at brighter magnitudes. It is possible that there is more than one structure, or maybe a set of structures, that conspire to make the North Galactic Cap appear asymmetric. It seems at least as likely that the spheroid itself is not symmetric about an axis through the center of the Milky Way and perpendicular to the Galactic plane. We have shown that our existing triaxial spheroid model is not perfect, but goes some way toward predicting the number of spheroid stars in the observations. We hope the model will improve as more large, coherent spheroid substructures are identified and as our fitting procedures evolve. 5. Conclusions The Sgr leading tidal tail is traced through the North Galactic Cap over the Center of the Milky Way, over S297+63-20.5, over the Sun, and at b = 30◦ it is heading toward the Galactic plane at a Galactic longitude of l = 205◦. We expect that it pierces the Galactic plane well outside the solar circle. We question whether some of the M giant stars that have been compared to Sgr dwarf disruption models are at the right distance to be members of the Sgr dwarf leading tidal tail. The main part of the Sgr dwarf leading tidal tail is not spatially coincident with the S297+63-20.5 overdensity in Virgo. The stars in S297+63-20.5 have line-of-sight, Galactic standard of rest velocities Vgsr = 130 ± 10 km s−1. The velocity dispersion is difficult to estimate because we do not know for sure which stars are part of the structure, but velocity dispersions of 10 km s−1 to 30 km s−1 are not unreasonable. Since the Sgr dwarf tidal tail would have negative velocities in this part of the sky, we determine that S297+63-20.5 is not associated with the leading tidal tail. Because the color of the turnoff, the density structure, and the Vgsr velocities do not match any known or expected Sgr debris, we suggest that it is instead part of a distinct merger -- 24 -- event. The Virgo Stellar Stream (VSS, Duffau et al. 2006) is at the same distance from the Sun as S297+63-20.5 and has a similar but not identical position in the sky and in Vgsr. It could be a the same as S297+63-20.5 or a structure within S297+63-20.5, but the relationship is not certain. The Virgo Overdensity (VOD, Juri´c et al. 2006) and S297+63-20.5 are in the same position in the sky and have a similar number of stars in excess of the smooth spheroid, but they are at a different distances from the Sun. Probably, there is a scale difference between our distance measurements and the VOD is the same as S297+63-20.5. We show that the number densities of F turnoff stars are not symmetric about the Galactic center at g0 ∼ 19.5, and that this discrepancy is not due to S297+63-20.5. Either the spheroid is asymmetric about the Galactic center, or there are additional substructures that conspire to be on the same side of the Galaxy as S297+63-20.5. Finally, we note that in every figure showing the spatial and velocity distribution in the spheroid, there are hints of extra unidentified substructure. In particular, we note an unexplained overdensity of BHB stars at (Λ⊙, g0) = (240◦, 16.7), and two additional moving groups near S297+63-20.5 with velocities Vgsr = −168±10kms−1 and Vgsr = −76±10kms−1. HJN acknowledges funding from the National Science Foundation (AST-0307571, AST- 0607618, AST-0612213), the NASA NY Space Grant, and John Huberty. TCB acknowledges partial funding for this work from National Science Foundation grants AST 04-06784 and PHY 02-16783: Physics Frontiers Center/Joint Institute for Nuclear Astrophysics (JINA). Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participating Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel, Cambridge University, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Ad- vanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for -- 25 -- Astrophysics (MPA), New Mexico State University, Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Obser- vatory, and the University of Washington. -- 26 -- REFERENCES Adelman-McCarthy, J., et al. 2007, ApJS, in press. Abazajian et al. 2003, AJ, 126,2081 Bell, E. F. et al. 2007, ApJ, submitted Bellazzini, M., Newberg, H. J., Correnti, M., Ferrar, F. R., & Monaco, L. 2006, a, 457, L21 Belokurov, V. et al. 2006a, ApJ 642, L137. Belokurov, V. et al. 2006b, ApJ 647, L111. Belokurov, V. et al. 2007a, ApJ 658,337 Belokurov, V. et al. 2007b, ApJ Letters, in press, astro-ph/0701790 Chou, M.-Y., Majewski, S. R., et al. 2006, ApJ, submitted, astro-ph/0605101 Dehnen, W. & Binney, J. J. 1998, MNRAS, 298, 387 Duffau, S., Zinn, R., Vivas, A. K., Carraro, G., Mendez, R. A., Winnick, R., & Gallart, C. 2006, ApJ, 636, L97 Fellhauer, M., et al. 2006, ApJ, 651, 167 Fukugita, M., Ichikawa,T., Gunn, J. E., Doi, M., Shimasaku, K., Schneider, D. P. 1996, AJ, 111, 1758 Gunn, J. E. et al. 1998, AJ, 116, 3040 Gunn, J. E. et al. 2006, AJ, 131, 2332 Helmi, A. et al. 2005, ApJ, 586, 195 Helmi, A. 2004, ApJ, 610, L97 Hogg, D. W., Finkbeiner, D. P., Schlegel, D. J., & Gunn, J. E. 2001, AJ, 122, 2129 Ibata, R. A., Irwin, M. J., Lewis, G. F., Ferguson, A. M. N., & Tanvir, N 2003, MNRAS 340, L21 Ivezi´c, Z., et al. 2004, Astronomische Nachrichten, 325, 583 Johnston, K. V., Law, D. R., & Majewski, S. R. 2005, ApJ, 619, 800 -- 27 -- Juric, M., et al. 2006, astro-ph/0510520 Law, D. R., Johnston, K. V. & Majewski, S. R. 2005, ApJ, 619, 807 Lenz, D. D., Newberg, H. J., Rosner, R., Richards, G. T., & Stoughton, C. 1998, ApJS, 119, 121 Majewski, S. R., Skrutskie, M. F., Weinberg, M. D., and Ostheimer, J. C. 2003, ApJ, 599, 1082 Martinez-Delgado, D., Penarrubia, J., Juric, M., Alfaro, E. J., Ivezic, Z. 2007, Ap. J. Suppl., 660, 1264 Martinez-Delgado, D., Gomez-Flechoso, M. A., Aparicio, A., Carrera, R. 2004, ApJ, 601,242 Moultaka, J., Ilovaisky, S. A., Prugniel, P., & Soubiran, C. 2004, PASP, 116, 693 Monaco, L., Bellazzini, M., Bonifacio, P., Buzzoni, A., Ferraro, F. R., Marconi, G., Sbordone, L., & Zaggia, S. 2006, A&A, 464, 201 Newberg, H., Yanny, B., et al. 2002, ApJ, 569, 245 Newberg, H., Yanny, B., et al. 2003, ApJ, 596, L191 Newberg, H., & Yanny, B. 2005, in ASP Conf. Ser. 338: Astrometry in the Age of the Next Generation of Large Telescopes, ed. P. K. Seidelmann & A. K. B. Monet, 210, astro-ph/0502386 Newberg, H., & Yanny, B. 2006, in JPC Conf. Ser.: Physics at the end of the Galactic Cosmic Ray Spectrum, ed. G. Thomson & P. Sokolsky, astro-ph/0507671 Odenkirchen, M. et al. 2006, AJ, 126, 2385 Parker, J. E., Humphreys, R. M., & Larsen, J. A. 2003, AJ, 126, 1346 Pier, J. R., Munn, J. A., Hindsley, R. B., Hennessy, G. S., Kent, S. M., Lupton, R. H., and Ivezi´c, Z. 2003, AJ, 125, 1559 Savage, C., Newberg, H. J., Freese, K., & Gondolo, P. 2006, Journal of Cosmology and Astroparticle Physics, 7, 3 Schlegel, D.J., Finkbeiner, D.P., & Davis, M. 1998, ApJ, 500, 525 Smith, J. A. et al. 2002, AJ, 123, 2121 -- 28 -- Stoughton, C., et al. 2001, AJ, 123, 485 Tucker, D., et al. 2006, Astronomische Nachrichten, 325, 583 Vivas, A. K. et al. 2001, ApJ, 554, L33 Xu, Y., Deng, L. C. & Hu, J. Y. 2006, MNRAS, 368, 1811 Xu, Y., Deng, L. C. & Hu, J. Y. 2007, MNRAS, in press Yanny, B., Newberg, H. J., et al. 2000, ApJ, 540, 825 Yanny, B., Newberg, H. J., et al. 2003, ApJ, 588, 841 York, D.G. et al. 2000, AJ, 120, 1579 Zinn, R., Vivas, A. K., Gallart, C. & Winnick, R. 2004, ASP Conf. Series, 327, 92 This preprint was prepared with the AAS LATEX macros v5.2. -- 29 -- Table 1. SDSS Stripes with Sgr Stream F turnoffs Stripe Number incl.a ◦ l ◦ 13 15 16 17 18 19 20 21 22 19e 20e 21e 22e 23e 7.5 12.5 15.0 17.5 20.0 22.5 25.0 27.5 30.0 22.5 25.0 27.5 30.0 32.5 308.7 277.1 245.4 233.2 221.7 215.9 212.1 209.2 207.2 229.1 217.2 209.0 204.0 201.8 b ◦ 70.2 73.9 69.0 65.9 57.2 51.0 46.7 41.0 33.7 75.6 72.9 67.5 62.4 53.2 g0 mag 22.00 21.85 21.60 21.52 21.35 21.12 21.12 21.02 20.68 21.75 21.45 21.25 21.65 21.40 db kpc µc ◦ d Λ⊙ ◦ 36 34 30 29 27 24 24 23 20 32 28 26 31 28 195 186 175 170 160 153 149 143 136 179 175 170 165 156 253 242 231 225 216 209 205 200 193 232 226 220 218 209 (a)Inclination of stripe relative to celestial equator, node is at α = 95◦. (b)Inferred distance of stars from the Sun, assuming Mg(F) = +4.2. (c)Angle along an SDSS "Stripe" which navigates an inclined great circle on the sky with a pole at (α, δ) = (95◦, 0◦). (d)Sagittarius plane azimuth (e)Indicates a secondary peak in the stripe, tracing a different density ridge. -- 30 -- Fig. 1. -- Spatial positions of BHB stars within 15 kpc of the Sgr dwarf orbital plane. Notice the arc of the leading tidal tail descending from 40 kpc above the Galactic plane and falling down on the solar position at XSgr,GC = −8.5 kpc. The overdensity near the Sun appears unexpectedly wide compared to the width of the leading tidal tail at apogalacticon. Part of the trailing tidal tail can be seen at (XSGR,GC, YSGR,GC) = (−80, −40) kpc. The Galactic plane is at YSgr,GC = 0; the Sgr orbital plane is as measured by Majewski et al. (2003) and with the Galaxy-centered X, Y, Z convention as described in Newberg et al. (2003). Fig. 2. -- g0 magnitude vs. Λ⊙ for BHB (left) and BS (right) stars within 15 kpc of the Sgr dwarf orbital plane. The A stars were separated into blue horizontal branch (BHB) and blue straggler (BS) stars using the color separation defined by Fig. 10 of Yanny et al. (2000). The Sgr leading and trailing tidal tails are marked as "Leading" and "Trailing," respectively. The leading tidal tail is evident in BHB stars from (Λ⊙, g0) = (290◦, 19.0) and sloping down toward the center of the diagram. If the leading tail came down on the Solar position from close to the North Galactic pole, it would come down at Λ⊙ = 256◦, which is marked in the diagram. Note that it instead passes this angle, heading towards the Galactic anticenter. The Sgr leading tidal tail is also evident in BS stars two magnitudes fainter; a small fraction of these BS stars are observed leaking into the BHB sample on the left panel. The trailing tidal tail is evident from (Λ⊙, g0) = (185◦, 20.3) and sloping down toward the center of the diagram. There is a very large falloff in the number of BHB's in the trailing tail at Λ⊙ = 195◦. Stars brighter than about 15.5 magnitudes are saturated in the SDSS survey, and therefore have less accurate magnitudes. It is unclear whether the bright stars are BHB stars out to 10 kpc, or BS stars closer than 4 kpc. In the left panel, there is a density peak of unknown origin centered at (Λ⊙, g0) = (240◦, 16.7). BHB stars within 0.2◦ of the four globular clusters M53, NGC 5053, NGC 4147, and NGC 5466 were discarded before making this diagram, which reduced but did not remove this density peak. Fig. 3. -- Separation of the BHB stars on each side of the Sagittarius orbital plane. We divide the BHB stars in the left panel of Fig. 2 along the plane of the Sagittarius dwarf orbit. The plane of the orbit of the dwarf is close to the Galactic (X, Z) plane, so most of the stars in the left panel have 180◦ < l < 360◦ and most of the stars in the right panel have 0◦ < l < 180◦. Stars from four globular clusters have been removed from the plot. Note that the peak at (Λ⊙, g0) = (240◦, 16.7) is not in the direction of S297+63-20.5, at (l, b) = (297◦, 63◦); it is on the wrong side of the Sgr orbital plane. The (Λ⊙, g0) = (240◦, 16.7) is low enough signal-to- noise that when we further subdivided the data into bins that were 5 kpc wide in ZSGR,GC it could not be discerned. Note also in this figure that the trailing Sgr stream is more evident in the right panel near apogalacticon (on the left edge) and then is more prominent in the left panel at brighter magnitudes. The trailing tidal tail is faint, but possibly extends as bright as 17th magnitude at Λ⊙ = 256◦. -- 31 -- Fig. 4. -- Position of S297+63-20.5 compared to other detections of the Sagittarius tidal tails. We reproduce Figures 3 and 4 from Newberg et al. (2003), which show the positions of the A-colored stars of the Sgr stream selected from eleven SDSS stripes (filled circles are leading tidal debris and open circles are trailing tidal debris), and the positions of 2MASS M giants from Fig. 11 of Majewski et al. (2003) (one point for each star). The larger open squares show the positions of ten new detections of the leading Sgr tidal tail. The smaller open squares show the positions of other bits of tidal debris detected in the same SDSS stripes, which is possibly due to superposition of young leading tidal debris and old trailing tidal debris (Fellhauer et al. 2006). The smaller squares trace debris on the opposite side of the Sgr orbital plane. The cross shows the position of S297+63-20.5, at (l, b, R) = (297◦, 63◦, 18 kpc), where R is the distance from the Sun. Note that although the center of S297+63-20.5 appears to be in the plane of the Sgr leading tidal tail, the distance does not match previous detections of the leading tidal tail. Fig. 5. -- F Star Polar Plot of the North Galactic Cap. The bottom panel shows stars with 0.2 < (g − r)0 < 0.3 and (u − g)0 > 0.4 in the magnitude range 20.0 < g0 < 21.0. The North Galactic Pole is in the center of the plot, and the outer circle is at b = 30◦. The distance between concentric circles of constant Galactic latitude was stretched to preserve solid angle per pixel. Darker areas of the diagram contain more F turnoff stars. All of the magnitudes were corrected using the reddening map of Schlegel, Finkbeiner, & Davis (1998) before selection. The dark areas at low latitude and toward the Galactic center represent the smooth portion of the Galactic spheroid. The dark line from (l, b) = (205◦, 25◦) to (l, b) = (255◦, 70◦) is the leading tail of the Sgr tidal stream, descending toward the Galactic plane as one moves to the left. The Sgr stream turnoff is fainter than g0 = 21 at the longitude of S297+63-20.5. A pair of square brackets encloses the primary area containing the S297+63-20.5 stellar excess. The top panel shows a subtraction of the pixels on the top half of the polar diagram, which does not contain any obvious tidal debris, from the lower half, with the assumption that the star counts are symmetric about l = 0◦, 180◦. One sees in the subtracted diagram that S297+63-20.5 peaks in this magnitude range near (l, b) = (300◦, 60◦); the density decreases for b < 64◦ as one moves down the center "outrigger" (SEGUE) extension of photometric data at l = 300◦. The Orphan Stream (Belokurov et al. 2007a) is visible on the edge of the data at l = 255◦. The tidal tails of the Pal 5 globular cluster (Odenkirchen et al. 2003) are near the edge of the data at l = 0◦. Fig. 6. -- Brighter and fainter F Star Polar Plots. SDSS data is extracted as in Figure 5, except we select closer (brighter, 19 < g0 < 20) F turnoff stars (upper panel) and more distant (fainter, 21 < g0 < 22) stars (lower panel). The Monoceros structure is apparent as a density enhancement toward the Galactic anti-center (165◦ < l < 225◦) in the upper panel. The Sagittarius stream and the S297+63-20.5 structures are prominent in the lower panel, -- 32 -- as is Palomar 5 at l = 0◦, b = 45◦. The brighter (top) panel shows stars approximately 9 to 14.5 kpc from the Sun. We do not see evidence for a separate VOD at this distance. There is, however, a general asymmetry in the number of spheroid stars around the Galactic center at l = 0. The fainter (bottom) panel shows stars approximately 23 to 36 kpc from the Sun. Since the Sgr stream at the anticenter is most prominent in Fig 5, it is between 14.5 and 23 kpc from the Sun as it passes below a Galactic longitude of b = 30◦. Fig. 7. -- Magnitude distribution of F Turnoff Stars in Virgo. We show histograms of F star (0.2 < g − r < 0.4) counts in mirror image 1.5◦ radius fields centered on: (l, b) = (288◦, 62◦), upper panel, light line, the VSS field; (l, b) = (72◦, 62◦), heavy line: the mirror image Galactic field at l′ = 360 − l. In the lower panel: (l, b) = (300◦, 55◦), light line, the S297+63-20.5 field; and (l, b) = (60◦, 55◦), heavy line, its mirror image field. The strong asymmetry in number counts between the quadrant IV fields over their quadrant I mirror images is apparent at magnitudes 18.5 < g0 < 22.5 (8 kpc < d < 45 kpc). The matching curves indicate the predicted triaxial halo model counts associated with each pointing. Note that neither a triaxial model alone nor a single halo stream alone can match all the strongly asymmetric data (see text). Fig. 8. -- Colors and magnitudes of SDSS F turnoff stars which have spectra in SEGUE. The smaller black dots show the colors and magnitudes of stellar objects in the part of the S297+63-20.5 structure observed with SEGUE plate 2707. The turnoff of S297+63-20.5 is visible in the enhanced density of field stars at g − r ∼ 0.26 and g0 ∼ 20.5. Larger symbols indicate stars with 0.1 < (g −r)0 < 0.4 and 19.4 < g0 < 20.5 for which spectra were obtained in the same 1.5 degree radius field targeted with plates 2568 and 2707. The symbol type indi- cates the SEGUE target selection category: BHB, F sub-dwarf and F/G dwarf candidates are variously-sized filled circles; squares indicate low metalicity candidates, and crosses indicate cool white dwarf candidates. The targets selected for spectroscopy were heavily weighted toward low metalicity (generally bluer) candidates, and do not representatively sample the color distribution of all objects in the field (black dots) when g − r0 < 0.2. Fig. 9. -- Galactic standard of rest velocities of F stars near the center of S297+63-20.5. We selected from SEGUE plates 2707 and 2568 all 91 of the stars with 0.2 < (g − r)0 < 0.4 and 19.4 < g0 < 20.5 that also had radial velocity errors of less than 20 km s−1. In the upper panels, stars from plate 2707 (l, b) = (300◦, 55◦) are subdivided into brighter (19.4 < g0 < 20, upper left), and fainter (g0 > 20) subsamples. The middle two panels show the same subsets for plate 2568 (l, b) = (288◦, 62◦). The lower panels sum the two panels above them, for these two fields separated by ∼ 9◦ degrees on the sky (2 − 3 kpc at the distances implied by these turnoff stars). In each panel, the solid line histogram represents a Gaussian distribution of halo stars centered on Vgsr = 0 with σ = 120 km s−1. The Gaussians are normalized so that -- 33 -- the area under the curve equals the number of spectra in that panel minus the number of stars above the Gaussian in bins that have more than a 2.5σ excess. The dotted line shows the limits for a 2.5σ excess in each bin. Any data peak above the dotted line represents a peak of greater than 2.5 σ significance. There are two significant peaks apparent in the lower panels, one at vgsr = −168 km s−1 for closer, brighter stars in the lower left panel, and a very prominent peak at vgsr = 130 km s−1 in the lower right panel (fainter stars) that is associated with the S297+63-20.5 structure at an implied distance of ∼ 18 kpc from the sun. Fig. 10. -- Galactic standard of rest velocities for F stars near the center of S297+63-20.5. 115 stars were selected and displayed as in Fig.9, except they are bluer, 0.1 < (g − r)0 < 0.2. There are two significant peaks: one at Vgsr ∼ 150 km s−1 in the left middle panel, and one at Vgsr = −76 km s−1 in the upper left panel. They are both present in the sum of these two panels at the lower left. The outgoing (positive Vgsr) stars are possibly related to S297+63-20.5. The origin of the incoming stars, which are also present at some level in the lower right panel, is less clear, but these stars have similar velocities to the stars which Law, Johnston, & Majewski (2005) fit to the Sgr leading tidal tail. -- 1 -- Boo Dwarf Sgr Leading Tail Sgr Trailing Tail Galactic Center -80 -60 -40 -20 0 20 40 -100 -80 -60 X = 0.998X + 0.066Y kpc -40 -20 SGR,GC 0 20 40 60 c p k Z 2 7 9 . 0 - Y 2 3 2 . 0 + X 5 1 0 . 0 - = C G , R G S Y 23 22 21 20 19 g 0 18 17 16 15 14 Trailing BS,Leading -- 2 -- Leading g 0 Unknown BHB stars Blue Straggler stars 160 180 200 220 240 260 280 300 180 200 220 240 260 280 300 Λo. Λo. 23 22 Trailing 21 20 19 g 0 18 17 16 15 14 -- 3 -- Trailing g 0 Leading Trailing? NGP Leading Unknown Z_SGR,GC < 0 Z_SGR,GC > 0 160 180 200 220 240 260 280 300 180 200 220 240 260 280 300 Λo. Λo. -- 4 -- -- 5 -- Fold over l=0-180 axis and subtract 105 75 b=30 135 45 b=60 b=70 15 345 165 195 225 315 Sgr Leading tail S297+63-20.5 l=255 285 -- 6 -- 105 75 b=30 135 b=60 b=70 165 195 225 Monoceros l=255 105 285 75 b=30 135 b=60 b=70 165 195 15 345 45 315 45 15 Pal 5 345 225 Sgr Leading tail 315 l=255 285 S297+63-20.5 -- 7 -- -- 8 -- -- 9 -- -- 10 --
astro-ph/0009215
1
0009
2000-09-13T22:16:11
The Formation of the Milky Way Disk
[ "astro-ph" ]
We present theoretical results on the galactic abundance gradients of several chemical species for the Milky Way disk, obtained using an improved version of the two-infall model of Chiappini, Matteucci, & Gratton (1997) that incorporates a more realistic model of the galactic halo and disk. This improved model provides a satisfactory fit to the elemental abundance gradients as inferred from the observations and also to other radial features of our galaxy (i.e., gas, star formation rate and star density profiles). We discuss the implications these results may have for theories of the formation of the Milky Way and make some predictions that could in principle be tested by future observations.
astro-ph
astro-ph
Galaxy Disks and Disk Galaxies ASP Conference Series, Vol. 3 × 108, 2000 J.G. Funes S.J., and E.M. Corsini, eds. THE FORMATION OF THE MILKY WAY DISK Cristina Chiappini Department of Astronomy, Columbia University, Mail Code 5247, Pupin Hall, 550 West 120th Street, New York, NY 10027 Francesca Matteucci Dipartimento di Astronomia, Universit`a di Trieste, Via G.B. Tiepolo 11, I-34131 Trieste, Italy Donatella Romano SISSA/ISAS, Via Beirut 2-4, I-34014 Trieste, Italy Abstract. We present theoretical results on the galactic abundance gra- dients of several chemical species for the Milky Way disk, obtained using an improved version of the two-infall model of Chiappini, Matteucci, & Gratton (1997) that incorporates a more realistic model of the galactic halo and disk. This improved model provides a satisfactory fit to the ele- mental abundance gradients as inferred from the observations and also to other radial features of our galaxy (i.e., gas, star formation rate and star density profiles). We discuss the implications these results may have for theories of the formation of the Milky Way and make some predictions that could in principle be tested by future observations. 1. Results and Discussion In this work we adopt a chemical evolution model (see Chiappini, Matteucci, & Romano 2000) that assumes two main accretion episodes for the formation of the Galaxy: the first one forming the halo and bulge in a short timescale followed by a second one that forms the thin-disk, with a timescale which is an increasing function of the Galactocentric distance (being of the order of 7 Gyrs at the solar neighborhood). The present model takes into account in more detail than previously the halo density distribution and explores the effects of a threshold density in the star formation process, both during the halo and disk phases. The model also includes the most recent nucleosynthesis prescriptions concerning supernovae of all types, novae and single stars dying as white dwarfs. In the comparison between model predictions and available data, we have focused our attention on abundance gradients as well as gas, star and star formation rate distributions along the disk, since this kind of model has already proven to be quite successful in reproducing the solar neighborhood characteristics. We suggest that the mechanism for the formation of the halo leaves heavy imprints on the chemical properties of the outer regions of the disk, whereas 1 2 Chiappini, Matteucci & Romano the evolution of the halo and the inner disk are almost completely disentangled. This is due to the fact that the halo and disk densities are comparable at large Galactocentric distances and therefore the gas lost from the halo can substan- tially contribute to build up the outer disk. We also show that the existence of a threshold density for the star formation rate, both in the halo and disk phase, is necessary to reproduce the majority of observational data in the solar vicinity and in the whole disk. In particular, a threshold in the star formation implies the occurrence of a gap in the star formation at the halo-disk transition phase, in agreement with recent data. Our main conclusions are: • Our best-model predicts gradients in good agreement with the observed ones in PNe, H II regions and open clusters. • The outer gradients are sensible to the halo evolution, in particular to the amount of halo gas which ends up into the disk. This result is not surprising since the halo density is comparable to that of the outer disk, whereas is negligible when compared to that of the inner disk. Therefore, the inner parts of the disk (R < R⊙) evolve independently from the halo evolution. • We predict that the abundance gradients along the Galactic disk must have increased with time. This is a direct consequence of the assumed "inside-out" scenario for the formation of the Galactic disk. Moreover, the gradients of different elements are predicted to be slightly different, owing to their differ- ent nucleosynthesis histories. In particular, Fe and N, which are produced on longer timescales than the α-elements, show steeper gradients. Unfortunately, the available observations cannot yet confirm or disprove this, because the pre- dicted differences are below the limit of detectability. • Our model guarantees a satisfactory fit not only to the elemental abundance gradients but it is also in good agreement with the observed radial profiles of the SFR, gas density and the number of stars in the disk. • Our best model suggests that the average < [α/Fe]> ratios in stars slightly decrease from 4 to 10 kpcs. This is due to the predominance of disk over halo stars in this distance range and to the fact that the "inside-out" scenario for the disk predicts a decrease of such ratios. On the other hand we predict a substantial increase (∼ 0.3 dex) of these ratios in the range 10 -- 18 kpcs, due to the predominance, in this region, of the halo over the disk stars. Finally, we conclude that a relatively short halo formation timescale (≃ 0.8 Gyr), in agreement with recent estimates for the age differences among Galactic globular clusters, coupled with an "inside-out" formation of the Galactic disk, where the innermost regions are assumed to have formed much faster than the outermost ones, represents, at the moment, the most likely explanation for the formation of the Milky Way. This scenario allows us to predict abundance gradi- ents and other radial properties of the Galactic disk in very good agreement with observations. More observations at large Galactocentric distances are needed to test our predictions. References Chiappini, C., Matteucci, F., & Gratton, R. 1997, ApJ, 477, 765 Chiappini, C., Matteucci, F., & Romano, D. 2000, ApJ, (submitted)
astro-ph/0605643
2
0605
2006-06-05T20:46:13
Diffraction-Based Sensitivity Analysis of Apodized Pupil Mapping Systems
[ "astro-ph" ]
Pupil mapping is a promising and unconventional new method for high contrast imaging being considered for terrestrial exoplanet searches. It employs two (or more) specially designed aspheric mirrors to create a high-contrast amplitude profile across the telescope pupil that does not appreciably attenuate amplitude. As such, it reaps significant benefits in light collecting efficiency and inner working angle, both critical parameters for terrestrial planet detection. While much has been published on various aspects of pupil mapping systems, the problem of sensitivity to wavefront aberrations remains an open question. In this paper, we present an efficient method for computing the sensitivity of a pupil mapped system to Zernike aberrations. We then use this method to study the sensitivity of a particular pupil mapping system and compare it to the concentric-ring shaped pupil coronagraph. In particular, we quantify how contrast and inner working angle degrade with increasing Zernike order and rms amplitude. These results have obvious ramifications for the stability requirements and overall design of a planet-finding observatory.
astro-ph
astro-ph
Diffraction-Based Sensitivity Analysis of Apodized Pupil Mapping Systems Ruslan Belikov Mechanical and Aerospace Engineering, Princeton University [email protected] N. Jeremy Kasdin Mechanical and Aerospace Engineering, Princeton University [email protected] Robert J. Vanderbei Operations Research and Financial Engineering, Princeton University [email protected] ABSTRACT Pupil mapping is a promising and unconventional new method for high con- trast imaging being considered for terrestrial exoplanet searches. It employs two (or more) specially designed aspheric mirrors to create a high-contrast amplitude profile across the telescope pupil that does not appreciably attenuate amplitude. As such, it reaps significant benefits in light collecting efficiency and inner work- ing angle, both critical parameters for terrestrial planet detection. While much has been published on various aspects of pupil mapping systems, the problem of sensitivity to wavefront aberrations remains an open question. In this paper, we present an efficient method for computing the sensitivity of a pupil mapped sys- tem to Zernike aberrations. We then use this method to study the sensitivity of a particular pupil mapping system and compare it to the concentric-ring shaped pupil coronagraph. In particular, we quantify how contrast and inner working angle degrade with increasing Zernike order and rms amplitude. These results have obvious ramifications for the stability requirements and overall design of a planet-finding observatory. Subject headings: Extrasolar planets, coronagraphy, Fresnel propagation, diffrac- tion analysis, point spread function, pupil mapping, apodization, PIAA -- 2 -- 1. Introduction The impressive discoveries of large extrasolar planets over the past decade have inspired widespread interest in finding and directly imaging Earth-like planets in the habitable zones of nearby stars. In fact, NASA has plans to launch two space telescopes to accomplish this, the Terrestrial Planet Finder Coronagraph (TPF-C) and the Terrestrial Planet Finder In- terferometer (TPF-I), while the European Space Agency is planning a similar multi-satellite mission called Darwin. These missions are currently in the concept study phase. In addition, numerous ground-based searches are proceeding using both coronagraphic and interferomet- ric approaches. Direct imaging of Earth-like extrasolar planets in the habitable zones of Sun-like stars poses an extremely challenging problem in high-contrast imaging. Such a star will shine 1010 times more brightly than the planet. And, if we assume that the star-planet system is 10 parsecs from us, the maximum separation between the star and the planet will be roughly 0.1 arcseconds. Design Concepts for TPF-C. For TPF-C, for example, the current baseline design involves a traditional Lyot coronagraph consisting of a modern 8th-order occulting mask (see, e.g., Kuchner et al. (2004)) attached to the back end of a Ritchey-Chretien telescope having an 8m by 3.5m elliptical primary mirror. Alternative innovative back-end designs still being considered include shaped pupils (see, e.g., Kasdin et al. (2003) and Vanderbei et al. (2004)), a visible nuller (see, e.g., Shao et al. (2004)) and pupil mapping (see, e.g., Guyon (2003) where this technique is called phase-induced amplitude apodization or PIAA). By pupil mapping we mean a system of two lenses, or mirrors, that takes a flat input field at the entrance pupil and produce an output field that is amplitude modified but still flat in phase (at least for on-axis sources). The Pupil Mapping Concept. The pupil mapping concept has received considerable attention recently because of its high throughput and small effective inner working angle (IWA). These benefits could potentially permit more observations over the mission lifetime, or conversely, a smaller and cheaper overall telescope. As a result, there have been numerous studies over the past few years to examine the performance of pupil mapping systems. In particular, Guyon (2003); Traub and Vanderbei (2003); Vanderbei and Traub (2005); Guyon et al. (2005) derived expressions for the optical surfaces using ray optics. However, these analyses made no attempt to provide a complete diffraction through a pupil mapping system. More recently, Vanderbei (2006) provided a detailed diffraction analysis. Unfortunately, this analysis showed that a pupil mapping system, in its simplest and most elegant form, cannot achieve the required 10−10 contrast; the diffraction effects from the pupil mapping systems themselves are so detrimental that contrast is limited to 10−5. In Guyon et al. (2005) and -- 3 -- Pluzhnik et al. (2006), a hybrid pupil mapping system was proposed that combines the pupil mapping mirrors with a modest apodization of oversized entrance and exit pupils. This combination does indeed achieve the needed high-contrast point spread function (PSF). In this paper, we call such systems apodized pupil mapping systems. A second problem that must be addressed is the fact that a simple two-mirror (or two-lens) pupil mapping system introduces non-constant angular magnification for off-axis sources (such as a planet). In fact, the off-axis magnification for light passing through a small area of the exit pupil is directly proportional to the amplitude amplification in that small area. For systems in which the exit amplitude amplification is constant, the magnification is also constant. But, for high-contrast imaging, we are interested in amplitude profiles that are far from constant. Hence, off-axis sources do not form images in a formal sense (the "images" are very distorted.) Guyon (2003) proposed an elegant solution to this problem. He suggested using this system merely as a mechanism for concentrating (on-axis) starlight in an image plane. He then proposed that an occulter be placed in the image plane to remove the starlight. All other light, such as the distorted off-axis planet light, would be allowed to pass through the image plane. On the back side would be a second, identical pupil mapping system (with the apodizers removed), that would "umap" the off-axis beam and thus remove the distortions introduced by the first system (except for some beam walk -- see Vanderbei and Traub (2005)). Sensitivity Analysis. What remains to be answered is how apodized pupil mapping behaves in the presence of optical aberrations. It is essential that contrast be maintained during an observation, which might take hours during which the wavefront will undoubt- edly suffer aberration due to the small dynamic perturbations of the primary mirror. An understanding of this sensitivity is critical to the design of TPF-C or any other observatory. In Green et al. (2004), a detailed sensitivity analysis is given for shaped pupils and vari- ous Lyot coronagraphs (including the 8th-order image plane mask introduced in Kuchner et al. (2004)). Both of these design approaches achieve the needed sensitivity for a realizable mission. So far, however, no comparable study has been done for apodized pupil mapping. One obstacle to such a study is the considerable computing power required to do a full 2-D diffraction simulation. Aberrations Given by Zernike Polynomials. In this paper, we present an efficient method for computing the effects of wavefront aberrations on apodized pupil mapping. We begin with a brief review of the design of apodized pupil mapping systems in Section 2. We then present in Section 3 a semi-analytical approach to computing the PSF of systems such as pupil-mapping and concentric rings in the presence of aberrations represented by Zernike polynomials. For such aberrations, it is possible to integrate analytically the integral over -- 4 -- azimuthal angle, thereby reducing the computational problem from a double integral to a single one, eliminating the need for massive computing power. In Section 4, we present the sensitivity results for an apodized pupil mapping system and a concentric ring shaped pupil coronagraph, and compare the results. 2. Review of Pupil Mapping and Apodization In this section, we review the apodized pupil mapping approach and introduce the specific system that we study in subsequent sections. It should be noted that this apodized pupil mapping design may not be the best possible. Rather, it is merely an example of such a system that achieves high contrast. Other examples can be found in the recent paper by Pluzhnik et al. (2006). Our aim in this paper is not to identify the best such system. Instead, our aim is to develop tools for carrying a full diffraction analysis of any apodized pupil mapping system. 2.1. Pupil Mapping via Ray Optics We begin by summarizing the ray-optics description of pure pupil mapping. An on-axis ray entering the first pupil at radius r from the center is to be mapped to radius r = R(r) at the exit pupil (see Figure 1). Optical elements at the two pupils ensure that the exit ray is parallel to the entering ray. The function R(r) is assumed to be positive and increasing or, sometimes, negative and decreasing. In either case, the function has an inverse that allows us to recapture r as a function of r: r = R(r). The purpose of pupil mapping is to create nontrivial amplitude profiles. An amplitude profile function A(r) specifies the ratio between the output amplitude at r to the input amplitude at r (in a pure pupil-mapping system the input amplitude is constant). Vanderbei and Traub (2005) showed that for any desired amplitude profile A(r) there is a pupil mapping function R(r) that achieves it (in a ray-optics sense). Specifically, the pupil mapping is given by r R(r) = ± vuuut Z 0 2A2(s)sds. (1) Furthermore, if we consider the case of a pair of lenses that are planar on their outward-facing surfaces, then the inward-facing surface profiles, h(r) and h(r), that are required to obtain the desired pupil mapping are given by the solutions to the following ordinary differential -- 5 -- equations: and ∂h ∂r (r) = ∂h ∂r (r) = r − R(r) n − 1qz2 + n+1 R(r) − r n − 1q z2 + n+1 , , h(0) = z, h(0) = 0. (2) (3) n−1(r − R(r))2 n−1(R(r) − r)2 Here, n 6= 1 is the refractive index and z is the distance separating the centers (r = 0, r = 0) of the two lenses. Let S(r, r) denote the distance between a point on the first lens surface r units from the center and the corresponding point on the second lens surface r units from its center. Up to an additive constant, the optical path length of a ray that exits at radius r after entering at radius r = R(r) is given by Q0(r) = S(R(r), r) + n(h(r) − h(R(r))). (4) Vanderbei and Traub (2005) showed that, for an on-axis source, Q0(r) is constant and equal to −(n − 1)z.1 2.2. High-Contrast Amplitude Profiles If we assume that a collimated beam with amplitude profile A(r) such as one obtains as the output of a pupil mapping system is passed into an ideal imaging system with focal length f , the electric field E(ρ) at the image plane is given by the Fourier transform of A(r): E(ξ, η) = eπi ξ2 +η2 λf E0 λif Z −∞ −∞ ∞ ∞ Z e−2πi xξ+ yη λf A(px2 + y2)dydx. (5) Here, E0 is the input amplitude which, unless otherwise noted, we take to be unity. Since the optics are azimuthally symmetric, it is convenient to use polar coordinates. The amplitude profile A is a function of r = px2 + y2 and the image-plane electric field depends only on 1For a pair of mirrors, put n = −1. In that case, z < 0 as the first mirror is "below" the second. -- 6 -- image-plane radius ρ = pξ2 + η2: E(ρ) = = eπi ξ2 +η2 λf eπi ξ2 +η2 λf 1 λif 2π λif e−2πi rρ λf cos(θ−φ)A(r)rdθdr 2π Z 0 J0(cid:18)−2π rρ λf(cid:19) A(r)rdr. ∞ 0 ∞ Z Z 0 The point-spread function (PSF) is the square of the electric field: Psf(ρ) = E(ρ)2. (6) (7) (8) For the purpose of terrestrial planet finding, it is important to construct an amplitude profile for which the PSF at small nonzero angles is ten orders of magnitude reduced from its value at zero. A paper by Vanderbei et al. (2003a) explains how these functions are computed as solutions to certain optimization problems. The high-contrast amplitude profile used in the rest of this paper is shown in Figure 2. 2.3. Apodized Pupil Mapping Systems Vanderbei (2006) showed that pure pupil mapping systems designed for contrast of 10−10 actually achieve much less than this due to harmful diffraction effects that are not captured by the simple ray tracing analysis outlined in the previous section. For most systems of practical real-world interest (i.e., systems with apertures of a few inches and designed for visible light), contrast is limited to about 10−5. Vanderbei (2006) considered certain hybrid designs that improve on this level of performance but none of the hybrid designs presented there completely overcame this diffraction-induced contrast degradation. In this section, we describe an apodized pupil mapping system that is somewhat more complicated than the designs presented in Vanderbei (2006). This hybrid design, based on ideas proposed by Olivier Guyon and Eugene Pluzhnik (see Pluzhnik et al. (2006)), involves three additional components. They are 1. a preapodizer A0 to soften the edge of the first lens/mirror so as to minimize diffraction effects caused by hard edges, 2. a postapodizer to smooth out low spatial frequency ripples produced by diffraction effects induced by the pupil mapping system itself, and -- 7 -- 3. a backend phase shifter to smooth out low spatial frequency ripples in phase. Note that the backend phase shifter can be built into the second lens/mirror. There are several choices for the preapodizer. For this paper, we choose Eqs. (3) and (4) in Pluzhnik et al. (2006) for our pre-apodizer: A0(r) = A(r)(1 + β) A(r) + βAmax , where Amax denotes the maximum value of A(r) and β is a scalar parameter, which we take to be 0.1. It is easy to see that • A(r)/Amax ≤ A0(r) ≤ 1 for all r, • A0(r) approaches 1 as A(r) approaches Amax, and • A0(r) approaches 0 as A(r) approaches 0. Incorporating a post-apodizer introduces a degree of freedom that is lacking in a pure pupil mapping system. Namely, it is possible to design the pupil mapping system based on an arbitrary amplitude profile and then convert this profile to a high-contrast profile via an appropriate choice of backend apodizer. We have found that a simple Gaussian amplitude profile that approximately matches a high-contrast profile works very well. Specifically, we used Apupmap(r) = 3.35e−22(r/a)2 , where a denotes the radius of the second lens/mirror. The backend apodization is computed by taking the actual output amplitude profile as computed by a careful diffraction analysis, smoothing it by convolution with a Gaussian distribution, and then apodizing according to the ratio of the desired high-contrast amplitude profile A(r) divided by the smoothed output profile. Of course, since a true apodization can never intensify a beam, this ratio must be further scaled down so that it is nowhere greater than unity. The Gaussian convolution kernel we used has mean zero and standard deviation a/√100, 000. The backend phase modification is computed by a similar smoothing operation applied to the output phase profile. Of course, the smoothed output phase profile (measured in radians) must be converted to a surface profile (having units of length). This conversion requires us to assume a certain specific wavelength. As a consequence, the resulting design is correct only at one wavelength. The ability of the system to achieve high contrast degrades as one moves away from the design wavelength. -- 8 -- 2.4. Star Occulter and Reversed System It is important to note that the PSFs in Figure 2 correspond to a bright on-axis source (i.e., a star). Off-axis sources, such as faint planets, undergo two effects in a pupil mapping system that differ from the response of a conventional imaging system: an effective magni- fication and a distortion. These are explained in detail in Vanderbei and Traub (2005) and Traub and Vanderbei (2003). The magnification, in particular, is due to an overall narrowing of the exit pupil as compared to the entrance pupil. It is this magnification that provides pupil mapped systems their smaller effective inner working angle. The techniques in Section 3 will allow us to compute the exact off-axis diffraction pattern of an apodized pupil mapped coronagraph and thus to see these effects. While the effective magnification of a pupil mapping system results in an inner working angle advantage of about a factor of two, it does not produce high-quaity diffraction limited images of off-axis sources because of the distortion inherent in the system. Guyon (2003) proposed the following solution to this problem. He suggested using this system merely as a mechanism for concentrating (on-axis) starlight in an image plane. He then proposed that an occulter be placed in the image plane to remove the starlight. All other light, such as the distorted off-axis planet light, would be allowed to pass through the image plane. On the back side would be a second, identical pupil mapping system (with the apodizers removed), that would "umap" the off-axis beam and thus remove the distortions introduced by the first system (except for some beam walk -- see Vanderbei and Traub (2005)). A schematic of the full system (without the occulter) is shown in Figure 3. Note that we have spaced the lenses one focal length from the flat sides of the two lenses. As noted in Vanderbei and Traub (2005), such a spacing guarantees that these two flat surfaces form a conjugate pair of pupils. 3. Diffraction Analysis In Vanderbei (2006), it was shown that a simple Fresnel analysis is inadequate for vali- dating the high-contrast imaging capabilities we seek. Hence, a more accurate approximation was presented. In this section, we give a similar but slightly different approximation that is just as effective for studying pupil mapping but is better suited to the full system we wish to analyze. -- 9 -- 3.1. Propagation of General Wavefronts The goal of this section is to derive an integral that describes how to propagate a scalar electric field from one plane perpendicular to the direction of propagation to another parallel plane positioned downstream of the first. We assume that the electric field passes through a lens at the first plane, then propagates through free space until reaching a second lens at the second plane through which it passes. In order to cover the apodized pupil mapping case discussed in the previous section, we allow both the entrance and exit fields to be apodized. Suppose that the input field at the first plane is Ein(x, y). Then the electric field at a particular point on the second plane can be well-approximated by superimposing the phase- shifted waves from each point across the entrance pupil (this is the well-known Huygens- Fresnel principle -- see, e.g., Section 8.2 in Born and Wolf (1999)). If we assume that the two lenses are given by radial "height" functions h(r) and h(r), then we can write the exit field as Eout(x, y) = Aout(r) ∞ ∞ −∞ −∞ Z Z 1 λiQ(x, y, x, y) e2πiQ(x,y,x,y)/λAin(r)Ein(x, y)dydx, (9) where Q(x, y, x, y) = q(x − x)2 + (y − y)2 + (h(r) − h(r))2 + n(Z − h(r) + h(r)) (10) is the optical path length, Z is the distance between the planar lens surfaces, Ain(r) denotes the input amplitude apodization at radius r, Aout(r) denotes the output amplitude apodiza- tion at radius r, and where, of course, we have used r and r as shorthands for the radii in the entrance and exit planes, respectively. As before, it is convenient to work in polar coordinates: Eout(r, θ) = Aout(r) where ∞ Z 0 2π Z 0 1 λiQ(r, r, θ − θ) e2πiQ(r,r,θ−θ)/λ)Ain(r)Ein(r, θ)rdθdr, (11) Q(r, r, θ) = qr2 − 2rr cos θ + r2 + (h(r) − h(r))2 + n(Z − h(r) + h(r)). (12) For numerical tractability, it is essential to make approximations so that the integral over θ can be carried out analytically, thereby reducing the double integral to a single one. To this end, we need to make an appropriate approximation to the square root term: S = qr2 − 2rr cos θ + r2 + (h(r) − h(r))2. (13) -- 10 -- A simple crude approximation is adequate for the 1/Q(r, r, θ − θ) amplitude-reduction factor in Eq. (11). We approximate this factor by the constant 1/Z. The Q(r, r, θ − θ) appearing in the exponential must, on the other hand, be treated with care. The classical Fresnel approximation is to replace S by the first two terms in a Taylor series expansion of the square root function about (h(r) − h(r))2. As we already mentioned, this approximation is too crude. It is critically important that the integrand be exactly correct when the pair (r, r) correspond to rays of ray optics. Here is a method that does this. First, we add and subtract S(r, r, 0) from Q(r, r, θ) in Eq. (11) to get Q(r, r, θ − θ) = S(r, r, θ − θ) − S(r, r, 0) + S(r, r, 0) + n(cid:16)h(r) − h(r)(cid:17) + S(r, r, 0) + n(cid:16)h(r) − h(r)(cid:17) + S(r, r, 0) + n(cid:16)h(r) − h(r)(cid:17) . S(r, r, θ − θ)2 − S(r, r, 0)2 S(r, r, θ − θ) + S(r, r, 0) rr − rr cos(θ − θ) (S(r, r, θ − θ) + S(r, r, 0))/2 = = (14) So far, these calculations are exact. The only approximation we now make is to replace S(r, r, θ− θ) in the denominator of Eq. (14) with S(r, r, 0) so that the denominator becomes just S(r, r, 0). Putting this all together, we get a new approximation, which we refer to as the S-Huygens approximation: Eout(r, θ) ≈ Aout(r) λiZ ∞ Z 0 K(r, r) 2π Z 0 e2πi(cid:16)− rr cos(θ− θ) S(r,r,0) (cid:17)/λEin(r, θ)dθ Ain(r)rdr, (15) where K(r, r) = e2πi( r r S(r,r,0) +S(r,r,0)+n(h(r)−h(r)))/λ (16) (note that we have dropped an exp(2πinZ/λ) factor since this factor is just a constant unit complex number which would disappear anyway at the end when we compute intensities). The only reason for making approximations to the Huygens-Fresnel integral (9) is to simplify the dependence on θ so that the integral over this variable can be carried out analytically. For example, if we now assume that the input field Ein(r, θ) does not depend on θ, then the inner integral can be evaluated explicitly and we get Eout(r, θ) ≈ 2πAout(r) λiZ ∞ Z 0 K(r, r)J0(cid:18) 2πrr λS(r, r, 0)(cid:19) Ein(r) Ain(r)rdr, (17) Removing the dependency on θ greatly simplifies computations because we only need to compute a 1D integral instead of 2D. In the next subsection we will show how to achieve similar reductions in cases where the dependence of Ein on θ takes a simple form. -- 11 -- Figure 4 shows plots characterizing the performance of an apodized pupil mapping system analyzed using the techniques described in this section. The specifications for this system are as follows. The designed-for wavelength is 632.8nm. The optical elements are assumed to be mirrors separated by 0.375m. The system is an on-axis system and we therefore make the non-physical assumption that the mirrors don't obstruct the beam. That is, the mirrors are invisible except when they are needed. The mirrors take as input a 0.025m on-axis beam and produce a 0.025m pupil-remapped exit beam. The second mirror is oversized by a factor of two; that is, its diameter is 0.050m. The postapodizer ensures that only the central half contributes to the exit beam. The first mirror is also oversized appropriately as shown in the upper-right subplot of Figure 4. After the second mirror, the exit beam is brought to a focus. The focal length is 2.5m. The lower-right subplot in Figure 4 shows the ideal PSF (in black) together with the achieved PSF at three wavelengths: at 70% (green), 100% (blue), and 130% (red) of the design wavelength. At the design wavelength, the achieved PSF matches the ideal PSF almost exactly. Note that there is minor degradation at the other two wavelengths mostly at low spatial frequencies. We end this section by pointing out that the S-Huygens approximation given by (15) is the basis for all subsequent analysis in this paper. It can be used to compute the propagation between every pair of consecutive components in apodized pupil mapping and concentric ring systems. It should be noted that the approximation does not reduce to the standard Fresnel or Fourier approximations even when considering such simple scenarios as free-space propagation of a plane wave or propagation from a pupil plane to an image plane. Even for these elementary situations, the S-Huygens approximation is superior to the usual textbook approximations. 3.2. Propagation of Azimuthal Harmonics In this section, we assume that E(r, θ) = E(r)einθ for some integer n. We refer to such a field as an nth-order azimuthal harmonic. We will show that an nth-order azimuthal har- monic will remain an nth-order azimuthal harmonic after propagating from the input plane to the output plane described in the previous section. Only the radial component E(r) changes, which enables the reduction of the computation from 2D to 1D. Arbitrary fields can also be propagated, by decomposing them into azimuthal harmonics and propagating each azimuthal harmonic separately. Computation is thus greatly simplified even for arbi- trary fields, especially for the case of fields which can be described by only a few azimuthal harmonics to a high precision, such as Zernike aberrations, which we consider in subsection 3.3. This improvement in computation efficiency is important, because a full 2D diffraction -- 12 -- simulation of an apodized pupil mapping system with the precision of greater than 1010 typically overwhelms the memory of a mainstream computer. By reducing the computation from 2D to 1D, however, the entire apodized pupil mapping system can be simulated with negligible memory requirements and takes only minutes. Theorem 1 Suppose that the input field in an optical system described by (15) is an nth- order azimuthal harmonic Ein(r, θ) = Ei(r)einθ for some integer n. Then the output field is also an nth-order azimuthal harmonic Eout(r, θ) = Eo(r)einθ with radial part given by Eo(r) = 2πin−1Aout(r) λZ ∞ Z 0 K(r, r)Ei(r)Jn(cid:18) 2πrr λS(r, r, 0)(cid:19) Ain(r)rdr. Proof. We start by substituting the azimuthal harmonic form of Ein into (15) and regroup- ing factors to get Eout(r, θ) = Aout(r) λiZ ∞ Z 0 K(r, r) 2π Z 0 e2πi(cid:16)− rr cos(θ− θ) S(r,r,0) (cid:17)/λEi(r)einθdθ Ain(r)rdr = Aout(r) λiZ einθ ∞ Z 0 K(r, r)Ei(r) 2π Z 0 e2πi(cid:16)− rr cos(θ− θ) S(r,r,0) (cid:17)/λein(θ−θ)dθ Ain(r)rdr. The result then follows from an explicit integration over the θ variable: Eout(r, θ) = 2πin−1Aout(r) λZ einθ ∞ Z 0 K(r, r)Ei(r)Jn(cid:18) 2πrr λS(r, r, 0)(cid:19) Ain(r)rdr (18) ✷ 3.3. Decomposition of Zernike Aberrations into Azimuthal Harmonics The theorem shows that the full 2D propagation of azimuthal harmonics can be com- puted efficiently by evaluating a 1D integral. However, suppose that the input field is not an azimuthal harmonic, but something more familiar, such as a (l, m)-th Zernike aberration: -- 13 -- Ein(r, θ) = eiǫZm l (r/a) cos(mθ), (19) where ǫ is a small number. (ǫ/2π and ǫ/π are the peak-to-valley phase variations across the aperture of radius a for m = 0 and m 6= 0, respectively.) Recall that the definition of the nth-order Bessel function is Jn(x) = 1 2πin 2π Z 0 eix cos θeinθdθ. (20) From this definition we see that ikJn(x) are simply the Fourier coefficients of eix cos(θ). Hence, the Fourier series for the complex exponential is given simply by the so-called Jacobi-Anger expansion ∞ eix cos θ = Xk=−∞ ikJk(x)eikθ. (21) The Zernike aberration can be decomposed into azimuthal harmonics using the Jacobi-Anger expansion: ∞ eiǫZm l (r/a) cos(mθ) = ikJk(ǫZ m l (r/a))eikmθ (22) (23) Xk=−∞ = J0(ǫZ m l (r/a)) + ∞ Xk=1 ikJk(ǫZ m l (r/a))eikmθ Note that Jk(x) ≈ 1 k! (cid:16) x 2(cid:17)k for 0 ≤ x ≪ 1. Hence, if we assume that ǫ ∼ 10−3, then the k'th term is of the order 10−3k. The field amplitude in the high-contrast region of the PSF will be dominated by the k = 1 term and be on the order of 10−3. If we drop terms of k = 3 and above, we are introducing an error on the order of 10−9 in amplitude. The error in intensity will be dominated by a cross-product of the k = 3 and the k = 1 term, or 10−12 across the dark region. So, in this case, Zernike aberrations can be more than adequately modeled using just 3 azimuthal harmonic terms. For ǫ ∼ 10−2, the number of terms goes up to 5 for an error tolerance of 10−12. In practice, even this small number of terms was actually found to be overly conservative. In order to compute the full 2D response for a given Zernike aberration, we simply decompose it into a few azimuthal harmonics, propagate them separately, and sum the results at the end. This method could also be applied to any arbitrary field. -- 14 -- 4. Simulations The entire 4-mirror apodized pupil mapping system can be modeled as the following sequence of 7 steps: 1. Propagate an input wavefront from the front (flat) surface of the first pupil mapping lens to the back (flat) surface of the second pupil mapping lens as described in Section 2.3. 2. Propagate forward a distance f . 3. Propagate through a positive lens with focal length f to a focal plane f units down- stream. 4. Multiply by star occulter. 5. Propagate through free-space a distance f then through a positive lens to recollimate the beam. 6. Propagate forward a distance f . 7. Propagate backwards through a pupil mapping system having the same parameters as the first one. A similar analysis can be carried out for a concentric ring shaped pupil system, or even a pure apodization system, as follows: 1. Choose Ain to represent either the concentric ring binary mask or some other az- imuthally symmetric apodization. 2. Choose h as appropriate for a focusing lens and let h ≡ 0. 3. Propagate through this system a distance f to the image plane. 4. Multiply by star occulter. The theorem can be applied to every propagation step, so that an azimuthal harmonic will remain an azimuthal harmonic throughout the entire system. Hence, our computation strategy was to decompose the input field into azimuthal harmonics, propagate each one separately through the entire system by repeated applications of the theorem, and sum them at the very end. -- 15 -- Figure 5 shows a cross section plot of the PSF as it appears at first focus and second focus in our apodized pupil mapping system (the first focus plot is indistinguishable from the case of ideal apodization or concentric ring shaped pupils). There are two plots for second focus: one with the occulter in place and one without it. Note that without the occulter, the PSF matches almost perfectly the usual Airy pattern. With the occulter, the on-axis light is suppressed by ten orders of magnitude. The electric field for a planet is just a slightly tilted and much fainter field than the field associated with the star. Hence, the methods presented here (specifically using the (1, 1)- Zernike, i.e. tilt) can be used to generate planet images. Some such scenarios are shown in Figure 6. The first row shows how an off-axis source, i.e. planet, looks at the first focus. As discussed earlier, at this focal plane off-axis sources do not form good images. This is clearly evident in this figure. The second row shows the planet as it appears at the second image plane, which is downstream from the reversed pupil mapping system. In this case, the off-axis source is mostly restored and the images begin to look like standard Airy patterns as the angle increases from about 2λ/D outward. Figure 7 shows corresponding cross sectional plots for the apodized pupil mapping system at the second focus. The third row in Figure 6 shows how a planet would appear at a focal plane of a concentric ring shaped pupil system. Figure 8 shows how the off-axis source is attenuated as a function of the angle from optical axis, for the case of our apodized pupil mapping system (at second focus) and the concentric ring coronagraph. For the case of apodized pupil mapping, the 50% point occurs at about 2.5λ/D. Figures 9 and 10 show the distortions/leakage from an on-axis source in the presence of various Zernike aberrations, for apodized pupil mapping and the concentric ring shaped pupil systems, respectively. The Zernike aberrations are assumed to be 1/100th wave rms. Figure 11 shows the corresponding cross-section sensitivity plots for both the apodized pupil mapping system and the concentric ring shaped pupil system. From this plot it is easy to see both the tighter inner working angle of apodized pupil mapping systems as well as their increased sensitivity to wavefront errors. Finally, Figure 12 demonstrates contrast degradation measured at three angles, 2, 4, and 8λ/D, as a function of severity of the Zernike wavefront error. The rms error is expressed in waves. 5. Conclusions We have presented an efficient method for calculating the diffraction of aberrations through optical systems such as apodized pupil mapping and shaped pupil coronagraphs. We -- 16 -- presented an example for both systems and computed their off-axis responses and aberration sensitivities. Figures 11 and 12 show that our particular apodized pupil mapping system is more sensitive to low order aberrations than the concentric ring masks. That is, contrast and IWA degrade more rapidly with increasing rms level of the aberrations. Thus, for a particular telescope, our pupil mapping system will achieve better throughput and inner working angle, but suffer greater aberration sensitivity. We note that there is a spectrum of apodized pupil mapping systems, out of which we selected but one example. The two extremes, pure apodization and pure pupil mapping, both have serious drawbacks. On the one end, pure apodization loses almost an order of magnitude in throughput and suffers from an unpleasantly large IWA. At the other extreme, pure pupil mapping fails to achieve the required high contrast due to diffraction effects. There are several points along this spectrum that are superior to the end points. We have focused on just one such point, which is similar to the design suggested by Guyon et al. (2005). We leave it to future work to determine if this is the best design point. For example, clearly one can improve the aberration sensitivity by relaxing the inner working angle and throughput requirements. Such analysis is beyond the scope of this paper, but we have provided here the tools to analyze the sensitivity of these kinds of designs. Acknowledgements. This research was partially performed for the Jet Propulsion Laboratory, California Institute of Technology, sponsored by the National Aeronautics and Space Administration as part of the TPF architecture studies and also under JPL subcontract number 1260535. The third author also received support from the ONR (N00014-05-1-0206). REFERENCES M. Born and E. Wolf. Principles of Optics. Cambridge University Press, New York, NY, 7th edition, 1999. P.S. Carney and G. Gbur. Optimal apodizations for finite apertures. Journal of the Optical Society of America A, 16(7):1638 -- 1640, 1999. A. Goncharov, M. Owner-Petersen, and D. Puryayev. Intrinsic apodization effect in a com- pact two-mirror system with a spherical primary mirror. Opt. Eng., 41(12):3111, 2002. J.J. Green, S.B. Shaklan, R.J. Vanderbei, and N.J. Kasdin. The sensitivity of shaped pupil coronagraphs to optical aberrations. In Proceedings of SPIE Conference on Astro- nomical Telescopes and Instrumentation, 5487, pages 1358 -- 1367, 2004. -- 17 -- O. Guyon. Phase-induced amplitude apodization of telescope pupils for extrasolar terrerstrial planet imaging. Astronomy and Astrophysics, 404:379 -- 387, 2003. O. Guyon, E.A. Pluzhnik, R. Galicher, R. Martinache, S.T. Ridgway, and R.A. Woodruff. Exoplanets imaging with a phase-induced amplitude apodization coronagraph -- i. prin- ciple. Astrophysical Journal, 622:744, 2005. J.A. Hoffnagle and C.M. Jefferson. Beam shaping with a plano-aspheric lens pair. Opt. Eng., 42(11):3090 -- 3099, 2003. N.J. Kasdin, R.J. Vanderbei, D.N. Spergel, and M.G. Littman. Extrasolar Planet Finding via Optimal Apodized and Shaped Pupil Coronagraphs. Astrophysical Journal, 582: 1147 -- 1161, 2003. M.J. Kuchner, J. Crepp, and J. Ge. Finding terrestrial planets using eighth-order image masks. Submitted to The Astrophysical Journal, 2004. (astro-ph/0411077). E.A. Pluzhnik, O. Guyon, S.T. Ridgway, R. Martinache, R.A. Woodruff, C. Blain, and R. Galicher. Exoplanets imaging with a phase-induced amplitude apodization coronagraph -- iii. hybrid approach: Optical design and diffraction analysis. Submitted to The Astrophysical Journal, 2006. (astro-ph/0512421). M. Shao, B.M. Levine, E. Serabyn, J.K. Wallace, and D.T. Liu. Visible nulling coronagraph. In Proceedings of SPIE Conference on Astronomical Telescopes and Instrumentation, number 61 in 5487, 2004. D. Slepian. Analytic solution of two apodization problems. Journal of the Optical Society of America, 55(9):1110 -- 1115, 1965. W.A. Traub and R.J. Vanderbei. Two-Mirror Apodization for High-Contrast Imaging. As- trophysical Journal, 599:695 -- 701, 2003. R. J. Vanderbei, N. J. Kasdin, and D. N. Spergel. Checkerboard-mask coronagraphs for high-contrast imaging. Astrophysical Journal, 615(1):555, 2004. R.J. Vanderbei. Diffraction analysis of 2-d pupil mapping for high-contrast imaging. Astro- physical Journal, 636:528, 2006. R.J. Vanderbei, D.N. Spergel, and N.J. Kasdin. Circularly Symmetric Apodization via Starshaped Masks. Astrophysical Journal, 599:686 -- 694, 2003a. R.J. Vanderbei, D.N. Spergel, and N.J. Kasdin. Spiderweb Masks for High Contrast Imaging. Astrophysical Journal, 590:593 -- 603, 2003b. -- 18 -- R.J. Vanderbei and W.A. Traub. Pupil Mapping in 2-D for High-Contrast Imaging. Astro- physical Journal, 626:1079 -- 1090, 2005. This preprint was prepared with the AAS LATEX macros v5.2. -- 19 -- h(cid:11)r(cid:12) r R(cid:11)(cid:97)r(cid:12) Z z 2 1.5 1 0.5 0 −0.5 −0.5 0 0.5 (cid:97)R(cid:11)r(cid:12) (cid:97)r (cid:97)h(cid:11)(cid:97)r(cid:12) (cid:19) Fig. 1. -- Pupil mapping via a pair of properly figured lenses. Light travels from top to bottom. 4 3.5 3 2.5 2 1.5 1 0.5 t u p n I o l t e v i t a e R e d u t i l p m A t u p t u O 0 −0.5 100 10−2 10−4 10−6 10−8 10−10 10−12 10−14 k a e P o t e v i t a e R y t i s n e t n I l Pupil−Plane Radius in fraction of Aperture 0 0.5 10−16 −60 −40 −20 0 20 40 60 Image−Plane Radius in L/D radians Fig. 2. -- Left. An amplitude profile providing contrast of 10−10 at tight inner working angles. Right. The corresponding on-axis point spread function. -- 20 -- Pre-apodize here Post-apodize here Occulter f f Pupil Mapper Reverse Pupil Mapper Fig. 3. -- The full pupil mapping system includes: a pair of lenses to shape the amplitude into a prolate-spheroidal-like amplitude profile, a focusing lens that concentrates the on- axis starlight into a small central lobe where a (not-depicted) occulter can block this light, followed by a recollimating lens and finally a reverse pupil mapping system that reforms the pupil with the starlight removed but any planet light (if present) intact. This final pupil is then fed a final focusing element (not shown) to form an image of off-axis sources. -- 21 -- 2nd Pupil Amplitude Map Target apodization Pre−apodizer Post−apodizer Achieved apodization 6 5 4 3 2 1 ) m ( t i h g e h r o r r i M / s n e L x 10−5 Lens/Mirror profiles First surface Second surface 0.005 0.01 0.015 0.02 Radius (m) 2nd Pupil Phase Map Phase smoothed Phase Ray−optics OPD 0.005 0.01 0.015 Radius (m) 0 0 0.005 0.01 0.015 0.02 Radius (m) PSFs ideal PSF achieved @ 632nm achieved @ 442nm achieved @ 822nm 5 10 working angle in units of λ/D 15 0 −5 −10 −15 k a e p o t e v i t a e r y t i s n e t n I l −20 0 t u p n I o t e v i t l a e R e d u t i l p m A t t u p u O 3.5 3 2.5 2 1.5 1 0.5 0 0 i s n a d a r n i e s a h P 0.2 0.15 0.1 0.05 0 −0.05 0 Fig. 4. -- Analysis of an apodized pupil mapping system using the S-Huygens approxima- tion with z = 15D and n = 1.5. Upper-left plot shows in red the target high-contrast amplitude profile and in blue the amplitude profile computed using the S-Huygens approx- imation through the apodized pupil mapping system. The other two curves depict the pre- and post-apodizers. Upper-right plot shows the lens profiles, red for the first lens and blue for the second. The lens profiles h and h were computed using a 5, 000 point discretization. Lower-left plot shows in red the computed optical path length Q0(r) and in blue the phase map computed using the S-Huygens propagation computed using a 5, 000 point discretiza- tion. Lower-right plot shows the PSF computed at three different wavelengths; the design value, 30% above that value, and 30% below it. -- 22 -- first focus (same as ideal PSF) second focus, no occulter second focus, with occultor 12 6 working angle in units of λ/D 10 8 14 16 18 0 −2 −4 −6 −8 −10 −12 −14 −16 −18 0 2 4 y t i s n e n t I 0 1 g o L Fig. 5. -- On-axis PSF at first focus (before occulter) and at second focus for cases of with and without occulter. Without the occulter, the second-focus PSF almost perfectly matches the usual Airy pattern. However, with the occulter, the second-focus on-axis PSF is suppressed by ten orders of magnitude. -- 23 -- Fig. 6. -- 2D pictures of planets for apodized pupil mapping and concentric rings. First row shows 2D intensity plots at first focus behind the occulter for planets at various angles relative to the on-axis star. Note that the system fails to form a clean image of the planets. Second row shows analogous plots at second focus. Note that the wavefront for the off-axis planet is mostly restored and the images begin to look like standard Airy patterns as the angle increases. Third row shows analogous plots for a concentric ring mask. -- 24 -- n o i t c u r t s b o o n h t i w e s n o p s e r o t e v i t l a e r y t i s n e t n I f o g o L 0 1 0 −0.5 −1 −1.5 −2 −2.5 −3 −3.5 −4 0 1 2 3 Tilt of 4 λ/D 2 λ/D 1 λ/D 7 8 9 10 4 5 working angle in units of λ/D 6 Fig. 7. -- Cross-sectional plots from the second row plots in Figure 6. Note that for angles of 3λ/D and above, the restored PSF looks very much like an Airy pattern with very little energy attenuation. However, as the angle decreases, the pattern begins to distort and the throughput begins to diminish. -- 25 -- Pupil mapping Concentric ring shaped pupil 2 3 working angle in units of λ/D 4 5 t u p h g u o r h t r e w o p l t a o T 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 1 Fig. 8. -- Off-axis source attenuation as a function of angle. Note that the 50% point occurs at about 2.5λ/D. -- 26 -- Fig. 9. -- Apodized pupil mapping sensitivities to the first nine Zernike aberrations. In each case, the rms error is 1/100th wave. The plots correspond to: Piston (0, 0), Tilt (1, 1), Defocus (2, 0), Astigmatism (2, 2) Coma (3, 1), Trefoil (3, 3), Spherical Aberration (4, 0), Astigmatism 2nd Order (4, 2), Tetrafoil (4, 4). -- 27 -- Fig. 10. -- Concentric-ring mask sensitivities to the first nine Zernike aberrations. -- 28 -- 0 −5 −10 0 −5 −10 0 −5 −10 0 (0,0) (1,1) Ideal Concentric Rings Pupil Mapping (2,0) (2,2) (3,1) (3,3) (4,0) (4,2) (4,4) 5 10 0 10 working angle in units of λ/D 5 0 5 10 t s a r t n o c f o 0 1 g o L Fig. 11. -- Radial profiles associated with the previous two Figures and overlayed one on the other. The green plots are for apodized pupil mapping whereas the blue plots are for the concentric ring mask. -- 29 -- Concentric rings, 4 λ/D Concentric rings, 8 λ/D Pupil mapping, 2 λ/D Pupil mapping, 4 λ/D Pupil mapping, 8 λ/D 10−5 10−10 10−4 (1,1) (2,0) 10−5 10−10 10−3 10−2 10−1 10−4 10−3 10−2 10−1 (2,2) (3,1) (3,3) 10−5 t s a r t n o C 10−10 10−5 10−10 10−5 10−10 10−4 10−3 10−2 10−1 10−4 10−3 10−2 10−1 10−4 10−3 10−2 10−1 (4,0) (4,2) (4,4) 10−5 10−10 10−5 10−10 10−5 10−10 10−4 10−3 10−2 10−1 10−4 10−1 Rms of aberration in units of wave 10−3 10−2 10−4 10−3 10−2 10−1 Fig. 12. -- Contrast degradation measured at three angles, 2, 4, and 8λ/D as a function of severity of the Zernike wavefront error measured in waves.
0707.0852
2
0707
2007-08-07T19:33:48
The CH2CN- molecule: Carrier of the lambda8037 diffuse interstellar band?
[ "astro-ph" ]
The hypothesis that the cyanomethyl anion CH2CN- is responsible for the relatively narrow diffuse interstellar band (DIB) at 8037.8 +- 0.15 Angstroms is examined with reference to new observational data. The 0_0^0 absorption band arising from the ^1B_1 - X ^1A' transition from the electronic ground state to the first dipole-bound state of the anion is calculated for a rotational temperature of 2.7 K using literature spectroscopic parameters and results in a rotational contour with a peak wavelength of 8037.78 Angstroms. By comparison with diffuse band and atomic line absorption spectra of eight heavily-reddened Galactic sightlines, CH2CN- is found to be a plausible carrier of the lambda8037 diffuse interstellar band provided the rotational contour is Doppler-broadened with a b parameter between 16 and 33 km/s that depends on the specific sightline. Convolution of the calculated CH2CN- transitions with the optical depth profile of interstellar Ti II results in a good match with the profile of the narrow lambda8037 DIB observed towards HD 183143, HD 168112 and Cyg OB2 8a. The rotational level populations may be influenced by nuclear spin statistics, resulting in the appearance of additional transitions from K_a = 1 of ortho CH2CN- near 8025 and 8050 Angstroms that are not seen in currently available interstellar spectra. For CH2CN- to be the carrier of the lambda8037 diffuse interstellar band, either a) there must be mechanisms that convert CH2CN- from the ortho to the para form, or b) the chemistry that forms CH2CN- must result in a population of K_a'' levels approaching a Boltzmann distribution near 3 K.
astro-ph
astro-ph
Astronomy&Astrophysicsmanuscript no. 7358manu October 26, 2018 © ESO 2018 The CH2CN− molecule: Carrier of the λ8037 diffuse interstellar band? Martin A. Cordiner1,2 and Peter J. Sarre1 1 School of Chemistry, The University of Nottingham, University Park, Nottingham, NG7 2RD, U.K. 2 Astrophysics Research Centre, School of Mathematics and Physics, Queen's University, Belfast, BT7 1NN, U.K. Received 23 February 2007 / Accepted 4 July 2007 ABSTRACT The hypothesis that the cyanomethyl anion CH2CN− is responsible for the relatively narrow diffuse interstellar band (DIB) at 8037.8 ± 0 absorption band arising from the 1B1 − X 1A′ transition from the 0.15 Å is examined with reference to new observational data. The 00 electronic ground state to the first dipole-bound state of the anion is calculated for a rotational temperature of 2.7 K using literature spectroscopic parameters and results in a rotational contour with a peak wavelength of 8037.78 Å. By comparison with diffuse band and atomic line absorption spectra of eight heavily-reddened Galactic sightlines, CH2CN− is found to be a plausible carrier of the λ8037 diffuse interstellar band provided the rotational contour is Doppler-broadened with a b parameter between 16 and 33 km s−1 that depends on the specific sightline. Convolution of the calculated CH2CN− transitions with the optical depth profile of interstellar Ti ii results in a good match with the profile of the narrow λ8037 DIB observed towards HD 183143, HD 168112 and Cyg OB2 8a. The rotational level populations may be influenced by nuclear spin statistics, resulting in the appearance of additional transitions from Ka = 1 of ortho CH2CN− near 8025 and 8050 Å that are not seen in currently available interstellar spectra. For CH2CN− to be the carrier of the λ8037 diffuse interstellar band, either a) there must be mechanisms that convert CH2CN− from the ortho to the para form, or b) the chemistry that forms CH2CN− must result in a population of K′′ a levels approaching a Boltzmann distribution near 3 K. Key words. Astrochemistry -- ISM: lines and bands -- ISM: molecules -- ISM: atoms -- ISM: clouds 1. Introduction The origin of the unidentified diffuse interstellar bands (DIBs) remains one of the greatest challenges in astronomical spec- troscopy. The subject has been reviewed by Herbig (1995) and Sarre (2006) who have highlighted research that points towards organic molecules as likely candidates for at least some of the more than 300 DIBs. Sarre (2000) has put forward a hypothe- sis that 'some, possibly many' of the diffuse interstellar bands arise from electronic transitions between ground and dipole- bound states of negatively charged polar molecules or small po- lar grains. It was noted that the rQ0(1) line of the origin band of the 1B1 − X 1A′ transition of CH2CN− occurs at 8037.8 Å, in good correspondence with the peak absorption wavelength of a diffuse interstellar band at 8037.9 ± 0.3 Å (Galazutdinov et al., 2000). 2. Molecular anions as diffuse band carriers 2.1. Previousstudies The electronic absorption spectrum of C− 7 in the visible region was studied by Tulej et al. (1998) and five optical absorption bands were reported to match with the wavelengths of known DIBs. Models of diffuse cloud chemistry by Ruffle et al. (1999) that incorporated the desorption of seed molecules from grain surfaces were able to reproduce large abundances of C− 7 (approx- imately 10−9nH), and also of C7H−, albeit only under a limited Send [email protected] offprint requests to: Martin Cordiner, e-mail: set of physical and chemical conditions and only for short peri- ods of time. The high electron affinity and density of vibrational states of C7 permits rapid radiative electron attachment such that anion formation has been calculated to occur every time an electron collides with C7 (Terzieva & Herbst, 2000). However, 7 as a DIB carrier was quashed when the early promise of C− high resolution observational and laboratory spectroscopy by Galazutdinov et al. (1999), Sarre & Kendall (2000), Lakin et al. (2000) and McCall et al. (2001) identified that the match be- tween the DIBs and the wavelengths, strengths and profiles of the C− 7 optical absorption bands was too poor to constitute an assignment. Based on laboratory spectroscopy it was suggested by Guthe et al. (2001) that CH2CC− might be a diffuse band car- rier but a detailed assessment by McCall et al. (2002b) showed that the wavelength match to the diffuse band at 6993 Å was not acceptable and commented that the non-observation of ad- ditional K sub-bands argued against an assignment to CH2CC−. The CH2CN− molecule has many spectroscopic characteristics in common with CH2CC−, both molecules being near-prolate asymmetric tops with ortho and para forms and both possessing dipole-bound excited states, but CH2CN− has a closed electronic shell. 2.2. AnionchemistryintheISM Chemistry in the diffuse ISM (with kinetic temperatures of ∼ 50 K, densities ∼ 100 cm−3 and strong UV radiation fields) has been shown to be rich and complex by observations which include the detection of polyatomic molecules such as H+ 3 (McCall et al., 1998), C3 (Maier et al., 2001), and HCO+, 2 M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? HCN, H2CO, C2H and c-C3H2 (Lucas & Liszt, 1996a,b, 2000; Liszt & Lucas, 2001) at high fractional abundances similar to those observed in dark clouds. Conventional models of gas- phase diffuse cloud chemistry struggle to explain how such high molecular abundances can be maintained in the interstellar UV radiation field. However, as discussed by Duley & Williams (1984) and Hall & Williams (1995), the availability of erosion products from carbonaceous dust grains, including polyynes (Cn chains) and small molecules such as diacetylene (C4H2), could provide a reservoir of reagents to fuel a complex network of or- ganic reactions in diffuse clouds and other photon-dominated re- gions. Large molecules should be able to resist photodissociation in the interstellar UV field by undergoing internal conversion in which the energy of absorbed photons is redistributed into the vibrational modes of the molecule due to the high density of states. To date the role of anions in interstellar reaction networks has generally been considered to be minimal although some con- sideration of their possible detection in the interstellar medium has been made (Sarre, 1980). No reaction pathways involving anions were present in the 'comprehensive' diffuse cloud mod- els of van Dishoeck & Black (1986) (where ∼ 500 reactions were modelled), and only a few atomic and diatomic anions are present in the UMIST 1999 database of ∼ 4000 astrochemical re- actions involving ∼ 400 different species (Le Teuff et al., 2000). This is at least partly due to the lack of observational evidence for anions in space, but also reflects a common view that anions should be rapidly photoionised in the interstellar UV field. The abundance of interstellar anions in relatively diffuse me- dia depends in large part on the balance between radiative elec- tron attachment (Dalgarno & McCray, 1973) and photodetach- ment by the interstellar radiation field. As discussed by Herbst (1981) and Herbst & Petrie (1997), a viable route for molecular anion formation in the ISM involves attachment of a free elec- tron via the formation of an excited temporary anion that sta- bilises through radiative transitions to the ground state. Provided the radiative transition rate is high, anions should be able to form at a sufficient rate that -- relative to their neutral counterparts -- appreciable anion abundances may occur depending on the local interstellar gas pressure and radiation field. Chemical models of the circumstellar envelope of the carbon star IRC+10°216 by Millar et al. (2000) predict observable abundances of Cn −, CnH− and CnN−. The discovery of the microwave signature of C6H− in IRC+10°216 and TMC-1 by McCarthy et al. (2006) at fractional abundances (relative to the neutral species C6H) of about 2.5% and 1%, respectively, proves that anions arise in some regions of space in significant quantities. 2.3. Dipole-boundstates Anions derived from strongly polar neutral parent molecules (or grains) can support one or more 'dipole-bound' electronic states at energies near to the detachment threshold and can po- tentially form the excited states in electronic absorption spec- tra in the visible region of the spectrum. Dipole-bound elec- tronic states were first predicted by Fermi & Teller (1947) who showed that a theoretical point dipole with a sufficiently large dipole moment (> 1.625 D) can bind an electron in a diffuse orbital. Theoretical studies (Crawford, 1970; Garrett, 1978; Gutsev & Adamowicz, 1995a,b,c) and experimental work (Moran et al., 1987; Brinkman et al., 1995; Lykke et al., 1987; Desfranc¸ois et al., 1996) has since demonstrated that in practice, dipole moments µ & 2.5 D are required for a molecule to sup- port at least one dipole-bound electronic state. Laboratory stud- ies of nitromethane (Compton et al., 1996) and of cyanoacety- Sp. type Target Cyg OB2 5 O7 Ie Cyg OB2 8a O6 I Cyg OB2 12 B2 I - B8 Ia HD 168112 HD 169034 HD 183143 HD 186745 HD 229196 λ Cyg ζ Peg O5 B5 Ia B7 Ia B8 Ia O5 B5 V B8.5 V V 9.2 9.0 11.4 8.6 8.2 6.9 7.1 8.6 4.6 3.4 EB−V 1.94 1.59 2.8 -- 3.4 1.01 1.3 1.24 0.96 1.22 0.04 0.00 d (pc) 700 1000 600 2000 1100 650 1600 1100 50 50 Table 1. Table showing sightline data for λ8037 observa- tions. For the EB−V values, B − V photometry are from Perryman & ESA (1997) and spectral types from the modal av- erages of all data referenced in the SIMBAD database (URL: http://simbad.u-strasbg.fr/sim-fid.pl); intrinsic stellar photome- try is from Wegner (1994) with the exception of Cyg OB2 5 and 8a for which photometry was taken from Massey & Thompson (1991) and Cyg OB2 12 which is of uncertain spectral type and photometry as discussed by Gredel & Munch (1994). Approximate stellar distances d (accurate to around ±50%), were calculated from the spectral types and absolute magnitudes of Wegner (2006) (stellar apparent magnitudes were dereddened assuming a ratio of visual to selective extinction (RV) of 3.1). lene and uracil (Sommerfeld, 2005) indicate that the presence of a near-threshold dipole-bound state can act as a 'doorway' for electron capture, and that strong electronic transitions occur from the dipole-bound state to the ground state, resulting in rapid radiative stabilisation of the anion. It is possible that this effect may enhance the electron capture rate for the strongly dipolar C6H molecule and therefore contribute towards the abundance of C6H− observed by McCarthy et al. (2006). 3. Spectroscopic observations and data reduction Medium resolution optical spectroscopic observations of early- type Galactic stars were performed using the High Resolution Echelle Spectrograph (HIRES) of the W. M. Keck Observatory, Hawaii by G. H. Herbig between 1995 and 1997 (private com- munication), who kindly made available the raw science expo- sures of eight heavily-reddened Galactic sightlines (shown in Table 1), including flat fields, Th/Ar and bias frames. Reductions were carried out using standard iraf echelle routines. In the re- gion of the λ8037 DIB special care was taken to eliminate tel- luric lines and fringing residuals by division with high S/N stan- dard star spectra. The S/N of the reduced spectra is typically be- tween 400 and 800 per pixel, with a resolving power of 42 500 (measured from the average of several unblended Th/Ar lines over the wavelength range of ∼ 7000 to 10 000 Å). In addition, near-UV spectra from ∼ 3200 to 4000 Å were obtained for the sight-lines towards Cyg OB2 8a, HD 168112 and HD 183143 and include spectra of a range of important interstellar species including Ca ii, Ti ii and various diatomic molecules that provide valuable information on the velocity distribution of the interstel- lar gases. The near-UV spectra typically have S/N ∼ 300 with a resolving power of 52 500. Around 8040 Å the wavelength drift between successive arcs was 0.02 Å, and for the near-UV ob- servations (around 3600 Å) the drift was 0.005 Å. Thus, abso- lute wavelength calibration errors should be ∆λ/λ . 2.5 × 10−6 (0.75 km s−1) for the red/NIR and . 1.4 × 10−6 (0.4 km s−1) for the near-UV spectra. M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? 3 4. Spectroscopy of CH2CN− 5. Results The neutral cyanomethyl radical CH2CN has a dipole mo- ment between c. 3.5 and 4.0 D (Ozeki et al., 2004) and an electron affinity of c. 1.55 eV (Lykke et al., 1987). The first dipole-bound state of the anion lies ≃ 10 meV below the ionisation continuum (Lykke et al., 1987; Moran et al., 1987; Gutsev & Adamowicz, 1995a). The CH2CN radical is planar but on attachment of an electron the hydrogen atoms move out of plane, resulting in a X 1A′ electronic ground state in Cs symme- try (Gutsev & Adamowicz, 1995a) with the mirror plane along the C-C-N backbone. In the excited 1B1 state, the dipole-bound electron is very diffuse with only a weakly perturbative inter- action with the rest of the molecule. Thus in the dipole-bound state the nuclear framework assumes a geometry which is al- most identical to that of the ground-state neutral radical. The 1B1 − X 1A′ transition is of perpendicular type (Lykke et al., 1987) with the transition dipole orientated along the c-axis and selection rules ∆J = ±1, ∆Ka = ±1 (Ka is used although it is not strictly a good quantum number). The rotational con- stants of CH2CN− were determined to very high accuracy by Lykke et al. (1987) using high-resolution fast-ion-beam autode- tachment spectroscopy. The rotational energy level structure and some examples of 1B1 − X 1A′ transitions are shown in Figure 4. of Lykke et al. (1987). Nuclear spin statistics dictate that there exist ortho (o) and para (p) forms of CH2CN− as occurs in the isoelec- tronic molecule cyanamide (NH2CN), discussed by Millen et al. (1962). If the ortho and para CH2CN− abundances reflect the o : p (3 : 1) nuclear spin degeneracies, transitions of ortho CH2CN− would be three times stronger than equivalent transi- tions of the para form. It not clear whether thermalisation of the rotational energy levels of CH2CN− would take place, but based on observations of other interstellar molecules, it is evi- dent that ortho-to-para abundance ratios do not necessarily fol- low the nuclear spin degeneracies. For example, according to Flower & Watt (1984), interconversion between ortho and para H2 occurs via collisional proton exchange in the reaction o-H2 + H+ −→ p-H2 + H+ + hν, which allows the population of the lowest J levels to reach thermal equilibrium with the ISM such that o : p is skewed away from the value of 3 : 1. In the case of CH2CN− the photodetachment (destruction) rate has been calcu- lated to be approximately 1.2×10−8 s−1 (E. Herbst & T. J. Millar, private communication); given the relatively slow rate of conver- sion between gas-phase ortho and para H2 by proton exchange with H+ (3 × 10−10nH+ s−1; Flower & Watt (1984)), it follows that in the neutral ISM where H+ densities are generally less than 0.1 cm−3, collisional exchange with gas-phase CH2CN− is unlikely to alter significantly the ortho-to-para population ratio. For molecules with equivalent H atoms the rotational level populations (and o : p ratios) can give clues as to the chemical reactions in which the molecule participates. For example, if a molecule is formed by reaction with H2, then the o : p ratio of the product may be influenced by the o : p ratio of the reagent H2 (see the case of c-C3H2, studied by Takakuwa et al. (2001)). Also, according to Dickens & Irvine (1999), the observed o : p ratio of H2CO, (which has the same nuclear spin statistical properties as CH2CN−), indicates that the molecules probably formed in thermal equilibrium with cold dust grains. Both of these types of reactions could potentially influence the o : p ratio of CH2CN−. Using the molecular constants of Lykke et al. (1987) and the asyrot fortran code (Birss & Ramsay, 1984), the 1B1 − X 1A′ transitions of CH2CN− were computed for a distribution of ro- tational level populations in equilibrium with the 2.74 K CMB. Isoelectronic with NH2CN (dipole moment 4.32 D; Tyler et al. (1972)), the cyanomethyl anion is expected to have a large dipole moment (∼ 4 D) such that in the diffuse ISM the radiative tran- sition rate is sufficient that collisional excitation of the molecule should be small. Small changes in the level of (J) rotational ex- citation of the molecule in fact have relatively little impact on the spectrum, which is dominated by the Q branch of the tran- sition at 8037.8 Å. Convolved with a Gaussian with Doppler b = 1 km s−1 and at a resolving power of 42 500, the calcu- lated spectrum is displayed in Figure 1 for comparison with the observed HIRES spectra. Examination of the spectrum of the standard star λ Cyg shows no evidence for significant stellar lines, fringing arti- facts or telluric residuals across the wavelength region. The most prominent features are the λ8026 DIB, which is among the narrowest known diffuse bands, and an absorption feature around 8037 Å. The λ8037 DIB is relatively narrow (FWHM ∼ 1.3 − 2 Å), but is overlapped by a broader absorption centered at around 8040 Å (noted by Herbig & Leka, 1991) that is most prominent towards Cyg OB2 12 and HD 183143. The λ8040 component has a FWHM of around 4 Å and forms an absorption peak at 8040.7 ± 0.3 Å. In the co-added (K i rest frame) spectra (shown in Figure 2), Gaussian fits to the peak of the λ8037 DIB give a wavelength of 8037.8 ± 0.15 Å for this diffuse interstellar band. The peak wavelength match of λ8037 with the 2.7 K 1B1 − X 1A′ CH2CN− spectrum shown in Figure 1 is very good. However, the computed spectrum clearly contains significant fine structure that is not present in any of the observed λ8037 profiles. The Q branch of the K′′ a = 0 transition creates the prominent peak at 8037.78 Å. The P branch is of rather low intensity compared to the R branch that produces the set of lines between 8035 and 8037.2 Å. Evidence for asymmetry of λ8037 can be seen in the blue degradation of the DIB spectra for the three Cyg OB2 sightlines and HD 183143, HD 186745 and HD 229196. The band shows significant evidence for profile variability across all of the sightlines, especially in comparing HD 186745 and HD 183143 with Cyg OB2. The variable strength and profile of λ8040 with respect to λ8037 suggests that it is probably caused by a different car- rier. Towards Cyg OB2 5 and HD 168112, the λ8040 DIB has FWHM ∼ 11.5 Å -- much broader than in the other sightlines -- and perhaps indicates the presence of another broad DIB. To isolate the λ8037 DIB for analysis is non-trivial among these other contaminating features. However, it is relatively narrow which assists its rejection in continuum fitting algorithms such that consistently repeatable continuum fits were possible. 5.1. OrthoandparaCH2CN− As detailed in Section 4, CH2CN− is expected to exist in the ISM in para and ortho forms, with the lowest occupied rotational lev- els being J′′ = 0, K′′ a = 1, respectively. The Ka = 1 ←− 0 transitions peak around 8037.8 Å, whereas the Ka = 0 ←− 1 and the Ka = 2 ←− 1 transitions peak at 8049.6 and 8024.8 Å respectively. The λ8037 co-added spec- trum for all eight sightlines is plotted in Figure 2 with para + a = 0 and J′′ = 1, K′′ 4 M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? y t i s n e t n i d e s i l a m r o N 1 0.96 0.92 0.88 0.84 0.8 0.76 0.72 0.68 0.64 − CH2CN (2.7 K) Cyg OB2 5 Cyg OB2 8a Cyg OB2 12 HD 168112 HD 169034 HD 183143 HD 186745 HD 229196 l Cyg 8025 8030 8035 8040 8045 8050 K I rest wavelength (Å) Fig. 1. Normalised (second order Chebyshev polynomial), telluric-corrected HIRES spectra of the region around 8037 Å. Spectra have been Doppler shifted to place the mean K i wave- length at rest (using λK i rest = 7698.9645 Å (Morton, 2003). λ Cyg is shown as an unreddened standard. Fitted continua are shown as dashed lines. The 2.74 K CH2CN− 1B1 − X 1A′ ori- gin band absorption spectrum is plotted at the top, calculated as- suming a single cloud at rest with Doppler b = 1 km s−1 and convolved to the resolving power R = 42 500 of the HIRES a = 0 transitions lie between 8035 and 8040 Å spectra. The K′′ and at this temperature the K′′ a = 1 transitions at around 8025 and 8046 Å are almost invisibly weak. The absorption feature at around 8026 Å is an unrelated diffuse interstellar band. ortho CH2CN− 1B1 − X 1A′ transitions, convolved with a rest cloud model (vLSR = 0 km s−1) with Doppler b = 18 km s−1 and an arbitrary intensity scaling. Two K′′ a population scenarios are plotted: (1) a 3 : 1 ratio of odd : even K′′ a level populations and (2) a 2.74 K Boltzmann distribution. Ka = 2‹ 1 Ka = 1‹ 0 Ka = 0‹ 1 y t i s n e t n i d e s i l a m r o N 1.06 1.04 1.02 1 0.98 0.96 0.94 8020 8025 8030 8035 8040 8045 8050 8055 K I rest wavelength (Å) a = 0 (para) and K′′ Fig. 2. Comparison of computed and observed (averaged) spec- tra in the 8037 A region. The bottom trace shows the co- added, normalised, telluric-corrected HIRES spectrum around λ8037 observed towards the stars listed in Table 1. Spectra were Doppler-shifted to place the weighted-mean K i wavelength at rest before co-addition. The 1B1 − X 1A′, {Ka = 2 ←− 1, Ka = 1 ←− 0, Ka = 0 ←− 1} modelled absorption spec- trum is shown, and includes the transitions arising from both a = 1 (ortho) CH2CN−. The middle spec- K′′ trum was calculated assuming a thermal Boltzmann distribution of Ka level populations. The upper spectrum shows the result taking the K′′ a populations to be determined solely by a 3 : 1 nuclear spin-statistical weight ortho : para ratio for odd : even a levels. The interstellar CH2CN− distribution is modelled as K′′ a single cloud at rest (v = 0 km s−1) with Gaussian line-shape (Doppler b = 18 km s−1), convolved to the resolving power R = 42 500 of the HIRES spectra. The respective peak absorp- tion wavelengths of the three Doppler-broadened Ka sub-bands are 8024.8, 8037.7, and 8049.6 Å, shown as dotted lines and cor- respond to ortho, para and ortho CH2CN− respectively. The predicted Ka = 2 ←− 1 transitions (with peak absorp- tion at ∼ 8024.8 Å) partially overlap the narrow λ8026 DIB. There is some evidence for a small blue shoulder on this DIB, at a wavelength slightly blue of the peak CH2CN− Ka = 2 ←− 1 absorption wavelength. The Ka = 0 ←− 1 feature (with peak ab- sorption at ∼ 8049.6 Å) falls close to a small absorption feature in the co-added spectrum. In this spectrum, the central depth of the λ8037 feature is 1.35 ± 0.16% and the central depth of the λ8049 feature is 0.3 ± 0.16% (2σ error estimates derived from the RMS noise of the continuum). The ratio of the central depth of the λ8037 to λ8049 features is consistent with a o : p ratio of ∼ 1 : 2.3, but given the mismatch between the calculated and ob- served profile and peak wavelength, it seems unlikely that λ8049 is caused by o-CH2CN−. From Figure 2, it is clear that if interstellar CH2CN− has a 3 : 1 'statistical' ratio of o : p states, it cannot be the carrier of the λ8037 DIB due to the lack of the ortho transitions at the ex- pected strengths in the observed spectra. The λ8026 DIB is too red relative to the Q branch (main peak) of the Ka = 2 ←− 1 M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? 5 y t i s n e t n i d e s i l a m r o N 1.005 1 0.995 0.99 0.985 1.005 1 0.995 0.99 0.985 1.005 1 0.995 0.99 0.985 1.005 1 0.995 0.99 Cyg OB2 5 HD 169034 Cyg OB2 8a HD 183143 Cyg OB2 12 HD 186745 HD 168112 HD 229196 −100 0 100 −100 0 100 LSR velocity (km s−1, l 0 = 8037.78 Å) 1.005 1 0.995 0.99 0.985 1.005 1 0.995 0.99 0.985 0.98 1.005 1 0.995 0.99 0.985 1 0.995 0.99 Fig. 3. Telluric-corrected, normalised DIB λ8037 spectra (thin black traces) and least-squares fitted CH2CN− 1B1 − X 1A′, Ka = 1 ←− 0 models (thick red traces). Models employ a sin- gle Gaussian interstellar cloud. Radial velocity shift, Doppler width, central depth and an additive continuum offset were free parameters in the fits. The DIB rest wavelength was set at λ0 = 8037.78 Å for the displayed velocity scale. Best-fitting model parameters are shown in Table 2 and the inferred CH2CN− in- terstellar velocity distributions plotted in Figure 4 (lines labeled 'Mdl'). vLSR 13.9 12.3 9.4 29.9 13.6 25.0 19.3 14.1 b 21.6 21.0 24.0 26.4 33.0 18.0 16.2 21.6 W8037 32.5 30.2 38.3 24.5 38.7 37.5 24.3 22.0 σc 0.0020 0.0024 0.0024 0.0018 0.0036 0.0022 0.0014 0.0020 Sightline Cyg OB2 5 Cyg OB2 8a Cyg OB2 12 HD 168112 HD 169034 HD 183143 HD 186745 HD 229196 Table 2. Least-squares fit parameters of the λ8037 CH2CN− models shown in Figure 3. The mean LSR velocity (vLSR / km s−1), Doppler cloud width (b / km s−1) and equivalent width of the model λ8037 DIB (W8037 / mÅ) are given along with the normalised continuum RMS of the observed spectra (σc), and the RMS of the model fit residuals (σ f ). σ f 0.0014 0.0016 0.0013 0.0021 0.0024 0.0012 0.0013 0.0012 transition to constitute a spectroscopic match with the model calculation. If however the Ka level populations have, or ap- proach, a Boltzmann distribution, then the transitions originating in Ka = 1 are sufficiently weak for CH2CN− to be a plausible carrier of the λ8037 diffuse interstellar band. Fig. 4. Each of the eight panels above shows LSR velocity- space apparent column density profiles for interstellar species for which data were available. Data for the K i 7699 Å, CH+ 3958 Å, Ca ii 3968 Å and Ti ii 3384 Å lines are shown. Hypothetical model CH2CN− velocity distributions (labeled 'Mdl') are shown at the top of each panel, calculated based on the fits shown in Figure 3 (parameters given in Table 2) with the assumption that the narrow λ8037 DIB is caused by the origin band 1B1 − X 1A′ transitions of CH2CN− at 2.74 K, and that the interstellar ve- locity distribution of the gas is Gaussian. The interstellar K i, CH+, Ca ii and Ti ii apparent column densities (Na) were calcu- lated from the normalised HIRES spectra (I/I0, where I0 is the continuum intensity) according to Na ∝ ln(I0/I) (see for exam- ple Savage & Sembach, 1991). For the right-hand panel, H i col- umn density profiles are from the Leiden/Argentine/Bonn (LAB) Survey of Galactic H i (Kalberla et al., 2005). 5.2. Modellingthe λ8037diffuseinterstellarband As shown by Figures 1 and 2, a large Doppler broadening of the interstellar gas is required for CH2CN− to be the carrier of λ8037. Based on the assumption that the λ8037 DIB is caused by the origin band 1B1 − X 1A′, K′′ a = 0 transitions of CH2CN− at 2.74 K, it is possible to use the shape of the DIB to in- fer the shape of the hypothetical interstellar CH2CN− velocity distribution. This distribution may then be compared with ob- 6 M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? LSR velocity (km s−1, l 0 −200 −100 0 = 8037.78 Å) 200 100 HD 183143 HD 168112 Cyg OB2 8a 1.005 1 0.995 0.99 0.985 0.98 0.975 0.97 1.005 1 0.995 0.99 0.985 1 0.995 0.99 0.985 0.98 0.975 0.97 y t i s n e n t i d e s i l a m r o N y t i s n e t n i d e s i l a m r o N y t i s n e t n i d e s i l a m r o N 8030 8035 8040 LSR wavelength (Å) 8045 Fig. 5. 2.74 K CH2CN− 1B1 − X 1A′, K′′ a = 0 transitions con- volved with the HIRES PSF and the interstellar Ti ii profiles (solid lines), and the K i profiles (dashed lines), observed to- wards HD 183143, HD 168112 and Cyg OB2 8a, overlaid on the normalised, telluric-corrected λ8037 DIB profiles (thin black histograms). served interstellar velocity profiles for different species in the observed sightlines. The computed 2.74 K absorption spectrum was subjected to a least-squares analysis whereby it was con- volved with a single Gaussian interstellar cloud model with the Doppler width, mean radial velocity and equivalent width floated as free parameters. A fourth free parameter allowed an additive shift in the continuum height if necessary. The least-squares pa- rameters for each sightline are shown in Table 2, and the corre- sponding DIB model spectrum and CH2CN− velocity distribu- tion are shown in Figures 3 and 4, respectively. The quality of the fits was monitored by comparing the normalised continuum RMS of the observed spectra (σc) with the normalised RMS of the model fit residuals (σ f ). 5.3. HypotheticalinterstellarCH2CN− distribution As shown by Figure 3, the model fits match the spectra to a high degree of accuracy for each of the eight sightlines. Variations in the width and structure of the DIB profile can be reproduced in each case, highlighted by the progression of the λ8037 FWHM from narrow to wide in the sequence HD 186745 → HD 183143 → Cyg OB2 12 → HD 169034. The progression corresponds to a sequential increase in the Doppler b of the cloud model: 16.2 → 18.0 → 24.0 → 33.0 km s−1. Such large b values would require unphysically large kinetic and/or turbulent gas velocities and clearly cannot arise in a single interstellar cloud, and this places some doubt on the validity of the single-cloud model used here. However, given the S/N of the spectra, there is no statistical justification for employing more than one cloud in the model. The Gaussian cloud model fits the DIB well as shown by σc ≈ σ f , implying that if CH2CN− is the λ8037 carrier, its interstellar velocity distribution must be approximately Gaussian. species available interstellar for five of survey (Kalberla et al., 2005, data Comparing the hypothetical CH2CN− velocity struc- lines in Figure 4 tures with the profiles of the calculated CH2CN− distributions are rea- shows that sonably close to the velocities of the diffuse interstellar clouds. Due to the lack of spectroscopic data for other interstellar the sightlines examined, H i data were obtained from the Leiden/Argentine/Bonn (LAB) at http://www.astro.uni-bonn.de/∼webrai/english/tools labsurvey.php). Although background H i emission contaminates these profiles, they at least indicate the likely velocity structure of a significant proportion of the neutral gas in these sightlines, aided by the fact that the stars are relatively distant (see Table 1). The peak of the calculated CH2CN− distribution generally lies within c. 10 -- 15 km s−1 of the peak H i column density. It should be noted that the mean LSR velocities of the foreground clouds traced by K i are positive for each sightline, which can be explained as a result of Galactic rotation and observational bias in the Galactic longitude of the targets. In Figure 4 there is an additional systematic small shift in the inferred CH2CN− distribution to a higher LSR velocity as compared to the K i. It is possible that this originates from the difficulty in defining the continuum of λ8037 as it sits atop a (variable) 8040 Å broader feature. Analysis of the velocity structure of interstellar gas in the ob- served sightlines has been performed to determine whether the DIB profiles can be reproduced by convolving the transitions of CH2CN− with the observed interstellar gas distributions. The re- sults of the least-squares fitting (Table 2) show that a Doppler b parameter of ∼ 20 km s−1 is required to achieve a good fit with the DIB. Typically, clouds traced by K i are much narrower M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? 7 than this (Welty & Hobbs, 2001), and even in heavily-reddened lines of sight such as those observed, the presence of multiple K i clouds separated in velocity space is insufficient to account for such broadening. Examining the data in left-hand panel of Figure 4, the HD 183143 least-squares model CH2CN− distribution shows gross similarities with the Ti ii and Ca ii profiles. It is ev- ident from these spectra that Ti ii has a broader and less-peaked distribution than the other species. In order of decreasing profile width and increasing structure, Ti ii comes first followed by Ca ii, then CH+, and finally K i. This sequence is believed to reflect the typical conditions in the clouds in which these species are generally found (see for example Jenkins, 1989; Crinklaw et al., 1994; Welty, 1998; Welty et al., 2003; Pan et al., 2005). Ti ii, due to its high condensation temperature tends to be found in the gas phase in hotter, more heavily shocked higher velocity clouds, and is depleted out onto grains in cool, quiescent diffuse clouds and molecular clouds. Ca has a similar condensation temperature to Ti and therefore comes out of the solid state to produce Ca ii in similar energetic/shocked regions to where Ti ii is observed, but will be ionised to Ca iii in more diffuse, strongly irradiated clouds, so it tends to be found in greatest abundances in warm yet moderately well-shielded regions. CH+ is photodissociated in less dense regions, but has been shown to have a broader velocity distribution than K i (Pan et al., 2005), consistent with the formation of this molecule in warmer haloes surrounding cold cloud cores (Crawford et al., 1994). K i is a good tracer for cool, quiescent, well-shielded clouds with low thermal broad- ening where the radiation field is attenuated and the photoioni- sation rate is relatively low. Based on this information and the breadth of the Ti ii distributions shown in Figure 4, gas traced by Ti ii may plausibly have a velocity distribution sufficient to provide the Doppler broadening required for CH2CN− to be the carrier of λ8037. Ti iii and other gases present in regions where H ii is found may exist over an even greater range of velocities. Figure 5 shows the convolution of the 2.74 K CH2CN− 1B1 − X 1A′, K′′ a = 0 transitions with the (HIRES) instrumen- tal PSF and with the interstellar Na i and Ti ii profiles for each of the three sightlines. Noiseless Na i and Ti ii velocity profiles were derived by modelling the Na i UV doublet and the Ti ii 3242 and 3384 Å lines using the vapid routine (Howarth et al., 2002). The accuracy of the derived Doppler b parameters was improved by the simultaneous modelling of two transitions of each species (originating in the same ground state), with differing oscillator strengths. Due to its similar chemistry, Na i traces the same type of interstellar material as K i, and was used in preference to the K i 7699 Å lines shown in Figure 4 due to the reduced satura- tion and telluric contamination of the Na i UV doublet. Atomic transition parameters were taken from Morton (2003). Within the spectral signal-to-noise, the profiles in Figure 5 calculated by convolution of the CH2CN− transitions with the interstellar Ti ii profiles compare favourably with the diffuse in- terstellar band wavelength and profile, especially for HD 168112 and HD 183143. Convolution by Ti ii gives a significantly better fit than convolution by Na i in all three cases, where the nar- rowness of the Na i Doppler spread produces structure in the re- sulting contour which is not seen in the observed DIB. These results demonstrate that the origin band CH2CN− 1B1 − X 1A′, Ka = 1 ←− 0 transitions at 2.74 K are capable of reproduc- ing the narrow λ8037 DIB towards HD 183143 and HD 168112 provided that the postulated CH2CN− carrier molecule is dis- tributed in velocity space in approximately the same way as in- terstellar Ti ii. Due to the lack of substructure in the observed λ8037 profile, it is very unlikely that CH2CN− could instead co- exist with Na i and K i in the cold neutral medium. For Cyg OB2 8a the peak absorption wavelength and profile match is reason- ably good for the case of both Na i and Ti ii, but there is clearly a significant level of structure in the calculated profiles that is not present in the observation. For the Ti ii-convolved calcula- tions, the relatively small differences between the calculated and observed profiles could plausibly be the result of Poisson noise of the spectrum or other features contaminating the DIB profile such as other interstellar features, telluric division or flat-fielding residuals. 6. Discussion Under the assumption that the CH2CN− distribution is the same as that of Ti ii -- with the implication that CH2CN− co-exists with gas-phase Ti ii -- the fit to the observed λ8037 spectrum towards HD 183143 and HD 168112 is very good (and to a lesser degree for Cyg OB2 8a). This result leads to two possibilities for the type of gas with which the postulated CH2CN− might be most closely associated: (1) The warm, shocked 'intercloud' medium; for the ∼ 50 interstellar clouds expected to be present in the heavily-reddened sightlines studied (see Welty et al., 2003), the warm, low density, shocked clouds with low depletions should dominate the Ti ii profiles (e.g. Crinklaw et al., 1994). (2) The neutral ISM; Ti ii is a good tracer of neutral hydrogen due to its ionisation potential of 13.6 eV (Stokes, 1978). Without knowl- edge of the precise H i velocity profiles, the possibility that the postulated CH2CN− co-exists with neutral atomic hydrogen gas cannot be ruled out. There are no reports of past attempts to detect CH2CN− or CH2CN in the diffuse ISM. The neutral radical CH2CN has been observed in the dense TMC-1 and Sgr B2 molecular clouds (see Irvine et al., 1988; Turner et al., 1990) through its pure rotational transitions. Fractional abundances are ∼ 10−9 to 10−11 relative to H2, and CH2CN was found to be similarly dis- tributed to CH3CN and C4H. The synthesis of CH2CN in the shocked diffuse interstellar clouds traced by Ti ii could be driven by the release of hydrocarbons into the gas phase during su- pernova shock-induced collisions between carbonaceous grains (Duley & Williams, 1984; Hall & Williams, 1995). Given a suf- ficient abundance of C2H+ C2H+ 4 + N −→ CH3CN+ + H, with a rate coefficient of ∼ 10−10 cm3 s−1 (Herbst & Leung, 1990), could proceed rapidly enough to produce sufficient quan- tities of CH3CN+ to allow dissociative recombination (DR) CH3CN+ + e −→ CH2CN + H 4 in the gas phase, the reaction (2) (1) to occur. This mechanism for CH2CN production is only vi- able if single H-atom loss is a significant channel in the DR of CH3CN+ (a process which is yet to be studied in detail), and if recombination occurs before CH3CN+ is destroyed by react- ing with neutral hydrogen. The dominant mechanisms by which CH2CN and CH3CN are expected to be destroyed in the diffuse ISM are by photodissociation and by reaction with C+. The most obvious pathway for production of the cyanomethyl anion CH2CN− is by radiative electron attachment to CH2CN: CH2CN + e −→ CH2CN− + hν, (3) though an alternative route exists via dissociative attachment to CH3CN: CH3CN + e −→ CH2CN− + H. (4) 8 M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? The last step was studied by Sailer et al. (2003) in the labora- tory. The dissociative electron attachment cross-section peaked at 4 × 10−23 m2 for electron energies of 3.2 eV, which provides an efficient route to the formation of CH2CN− from CH3CN. The cyanomethyl anion production channel is ∼ 500 times more efficient than those of the other possible anionic fragments (CHCN−, CCN−, CN− and CH3 −) for electrons energies between 1 and 10 eV. In the interstellar context, photon and cosmic ray impact on grains could provide a source of electrons with a few eV energy. If the dominant chemical reactions involving CH2CN− are assumed to be formation of the anion via radiative attachment (Equation (3)) and destruction via photodetachment, then in equilibrium, the anion-to-neutral ratio is given by [CH2CN−] [CH2CN] αne Γ (5) = where α is the rate of radiative electron attachment, ne is the electron density and Γ is the photodetachment rate. Following s-wave electron capture by CH2CN into an excited state of CH2CN− (Terzieva & Herbst, 2000), if subsequent radiative re- laxation occurs rapidly, either by vibrational transitions of the vi- brationally excited ground state or by electronic transitions from the dipole-bound state to the ground state, then α can be esti- mated using Equation (6) of Terzieva & Herbst (2000). In the diffuse ISM (at T = 100 K), this yields a value of α = 2.2 × 10−7 cm−3 s−1, assuming that the exited anion undergoes radiative stabilisation before the electron detaches back to the contin- uum. The photodetachment rate Γ has been calculated to be approximately 1.2 × 10−8 s−1 (E. Herbst & T. J. Millar, pri- vate communication) in the standard interstellar radiation field (Draine, 1978). Both parameters are uncertain due to the lack of laboratory or theoretical data for this molecule, but neverthe- less permit a rough estimate of the fraction of CH2CN likely to occur in anionic form. Using an electron density in the dif- fuse ISM of between 0.04 and 0.23 cm−3 (Welty et al., 2003), Equation (5) yields an anion-to-neutral ratio between 0.7 and 4.2. If CH2CN− is present in the warm neutral medium traced by Ti ii (and Ca ii) where T ∼ 4000 K (see Welty et al., 1996) then α = 3.4×10−8 cm−3 s−1 and the calculated anion-to-neutral ratio is 0.1 -- 0.7. A further test of the plausibility of CH2CN− as carrier of the λ8037 DIB is whether a feasible CH2CN abundance can be deduced from the equivalent width of the observed DIB. Polyatomic molecules are likely to be present in all of the heavily-reddened sightlines towards which λ8037 was observed. Cyg OB 12, Cyg OB2 5 and HD 183143 have been studied by McCall et al. (1998, 2002a), particularly with reference to interstellar H3 +, detected in all three sightlines with Doppler b ∼ 10 km s−1 which is substantially broader than the 13CO and HCO+ lines observed towards Cyg OB2 12. The HCO+ column density was found to be ∼ 1011 cm−2. Scappini et al. (2002) de- termined a likely total number density of n & 104 cm−3 for the gas containing the CO and HCO+, with the inference of dense, compact molecular clumps within a more diffuse medium. The approximate molecular hydrogen column density towards Cyg OB2 12 (from McCall et al., 2002a) is 6.5 × 1021 cm−2 such that the fractional HCO+ abundance is ∼ 1.5 × 10−11nH2. The equivalent width of the λ8037 DIB is 38.3 mÅ (see Table 2) which requires a column density of N(CH2CN−) = 6.7 × 1010 cm−2/ f where f is the oscillator strength of the tran- sition. Transitions between the ground and dipole-bound states are strongly observed in the lab (Lykke et al., 1987), suggest- ing that the value of f should be relatively large. Therefore, if f = 0.5 and assuming a CH2CN anion-to-neutral ratio of 1, a molecular fraction of CH2CN of 2 × 10−11nH2 is required to- wards Cyg OB2 12 to produce the observed λ8037 equivalent width. Thus, provided the 1B1 − X 1A′ transition is strong, the required interstellar CH2CN fractional column density is com- parable to that of HCO+ observed in the Cyg OB2 12 sightline. Averaged along the sightline, the required interstellar CH2CN fractional abundance is at the lower end of the range of val- ues found in dense molecular clouds (∼ 10−11 − 10−9nH2). If diffuse and dense cloud molecular abundances are similar, then enough CH2CN− can probably be produced to create the λ8037 DIB along heavily-reddened sightlines. The spectra observed by McCall et al. (2002a) show clearly that the interstellar material traced by H+ 3 has a broader veloc- ity distribution than the molecules (CO, C2, CN) which typi- cally trace dense molecular material. The H+ 3 line FWHM ob- served towards Cyg OB2 12, Cyg OB2 5 and HD 183143 range from 8 to 15 km s−1 whereas the CN line FWHM are all less than 4 km s−1. McCall et al. (2002a) conclude that 'the chem- istry that leads to H+ 3 is completely decoupled from that which is responsible for these heavier diatomics'. These factors lend support to the possibility that other molecules may be present in significant abundances in the more diffuse regions surrounding conventional molecular clouds. Very rapid production mechanisms are required for molecules such as CH2CN and its anion to exist in the warm diffuse regions traced by Ti ii, though PAHs and related large or- ganic species will survive for much longer in the UV radiation field with the potential to give rise to diffuse interstellar bands. The interstellar media containing Ti ii and Ca ii are largely co- spatial (see Crinklaw et al., 1994; Albert et al., 1993), and these refractory ions are observed predominantly in the warm neu- tral ISM. Welty et al. (1996) found the mean upper temperature limit of the clouds in their Ca ii survey to be 4100 K, though temperatures as low as a few hundred K were found in many cases. Titanium and calcium arise in the gas phase predomi- nantly due to grain destruction (Stokes, 1978), and it is plausible that molecules and DIB carriers may also arise in this way. There is no evidence in the literature for correlations between DIB strengths and Ti ii column densities, though Herbig (1993) found that the 5780 and 5797 DIBs do not correlate with the level of titanium depletion. This study suggests that further in- vestigation of a possible link between DIBs, molecules and Ti ii is warranted. 7. Conclusion Using the rotational constants of Lykke et al. (1987), the ab- sorption spectrum arising from the 1B1 − X 1A′ transition of CH2CN− has been computed for CH2CN− in equilibrium with the CMB at 2.74 K. The peak of the intrinsic absorption profile (at 8037.78 Å) is consistent with the peak of the diffuse interstel- lar absorption feature at 8037.8 ± 0.15 Å found in the spectra of eight heavily-reddened stars. If CH2CN− occurs in the ISM with the statistical ortho : para abundance ratio of 3 : 1 the 1B1− X 1A′ transitions with K′′ a = 1 would produce strong spectral features at wavelengths of approximately 8024.8 and 8049.6 Å that are not present in the observed interstellar spectra. However, if the a = 1 level is populated based on a Boltzmann distribution K′′ at 2.74 K, then the Ka = 0 ←− 1 and Ka = 2 ←− 1 sub- bands are sufficiently weak to be compatible with their absence from the observed spectra. A thermal (Boltzmann) distribution of K′′ a levels would be expected if an efficient mechanism exists M.A. Cordiner and P.J. Sarre: CH2CN−: Carrier of the λ8037 diffuse interstellar band? 9 for conversion between the ortho and para forms of CH2CN−. Alternatively, the formation mechanism(s) for CH2CN− may in- troduce a skewed (i.e. non 3 : 1) ratio of ortho and para forms. Higher signal-to-noise (& 2000) spectroscopic observations of the λ8037 region will assist in the search for the K′′ a = 1 sub- bands. Without detailed laboratory or theoretical studies, it is unclear whether or not the CH2CN− formation mechanism(s) would result in the required deviation of the ortho-to-para ratio from the statistical value of 3 : 1. The cyanomethyl anion CH2CN− is a plausible candidate for the carrier of the narrow diffuse interstellar band located at 8037.8 Å, provided a spectral line-broadening mechanism can be accounted for that produces an approximately Gaussian broad- ening corresponding to a Doppler b of between 16 and 33 km s−1 which corresponds with the observed Ti ii and Ca ii profiles to- wards HD 183143 and HD 168112, and is comparable to the H+ 3 linewidths towards Cyg OB2 12, Cyg OB2 5 and HD 183143. Measurement of the oscillator strength of the 1B1 − X 1A′ transition would enable the calculation of the interstellar col- umn density of CH2CN− that would be required to produce the λ8037 DIB. Assuming the transition is strong ( f ∼ 0.5), and the CH2CN anion-to-neutral ratio is large (∼ 1, as suggested by preliminary calculations of the anion formation and destruction rates), then towards Cyg OB2 12 the required average fractional abundance of neutral CH2CN relative to H2 is 2 × 10−11, which falls at the lower end of the observed range in dense clouds. The spectroscopic consistency between the λ8037 DIB and the calculated transitions of CH2CN− cannot be considered an assignment, but provides strong motivation for further study of this molecular anion including a search for the pure rotational transitions of interstellar CH2CN−. Acknowledgements. The authors would like to thank George Herbig for kindly making available extensive Keck HIRES data and Ian Howarth for providing the vapid software and support. MAC thanks EPSRC for a studentship and The University of Nottingham for financial support. We also thank Tom Millar and Steve Fossey for helpful comments, and a referee for suggested improvements to the manuscript. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France. The W. M. Keck Observatory is operated as a sci- entific partnership among the California Institute of Technology, the University of California, and NASA. The observatory was made possible by the generous financial support of the W. M. Keck Foundation. References Albert, C. E., Blades, J. C., Morton, D. C., et al. 1993, ApJS, 88, 81 Birss, F. W. & Ramsay, D. A. 1984, Computer Physics Communications, 38, 83 Brinkman, E. A., Berger, S., Marks, J., & Brauman, J. I. 1995, J. Chem. Phys., 99, 7586 Compton, R. N., Carman, H. S., Desfranc¸ois, C., Abdoul-Carime, H., & Scherman, J. P. 1996, J. Chem. Phys., 105, 6126 Crawford, I. A., Barlow, M. J., Diego, F., & Spyromilio, J. 1994, MNRAS, 266, 903 Crawford, O. H. 1970, Mol. Phys., 20, 585 Crinklaw, G., Federman, S. R., & Joseph, C. L. 1994, ApJ, 424, 748 Dalgarno, A. & McCray, R. A. 1973, ApJ, 181, 95 Desfranc¸ois, C., Abdoul-Carime, H., & Scherman, J. P. 1996, Int. J. Mod. Phys., 10, 1339 Dickens, J. E. & Irvine, W. M. 1999, ApJ, 518, 733 Draine, B. T. 1978, ApJS, 36, 595 Duley, W. W. & Williams, D. A. 1984, MNRAS, 211, 97 Fermi, E. & Teller, E. 1947, Phys. Rev., 72, 406 Flower, D. R. & Watt, G. D. 1984, MNRAS, 209, 25 Guthe, F., Tulej, M., Pachkov, M. V., & Maier, J. P. 2001, ApJ, 555, 466 Galazutdinov, G. A., Krełowski, J., & Musaev, F. A. 1999, MNRAS, 310, 1017 Galazutdinov, G. A., Musaev, F. A., Krełowski, J., & Walker, G. A. H. 2000, PASP, 112, 648 Garrett, W. R. 1978, J. Chem. Phys., 69, 2621 Gredel, R. & Munch, G. 1994, AA, 285, 640 Gutsev, G. L. & Adamowicz, L. 1995a, Chem. Phys. Lett., 246, 245 Gutsev, G. L. & Adamowicz, L. 1995b, J. Phys. Chem., 99, 13412 Gutsev, G. L. & Adamowicz, L. 1995c, Chem. Phys. Lett., 235, 377 Hall, P. & Williams, D. A. 1995, Ap&SS, 229, 49 Herbig, G. H. 1993, ApJ, 407, 142 Herbig, G. H. 1995, Annu. Rev. Astophys., 33, 19 Herbig, G. H. & Leka, K. D. 1991, ApJ, 382, 193 Herbst, E. 1981, Nature, 289, 656 Herbst, E. & Leung, C. M. 1990, A&A, 233, 177 Herbst, E. & Petrie, S. 1997, ApJ, 491, 210 Howarth, I. D., Price, R. J., Crawford, I. A., & Hawkins, I. 2002, MNRAS, 335, 267 Irvine, W. M., Friberg, P., Hjalmarson, A., et al. 1988, ApJ, 334, L107 Jenkins, E. 1989, in IAU Symp. 135: Interstellar Dust, 23 Kalberla, P. M. W., Burton, W. B., Hartmann, D., et al. 2005, VizieR Online Data Catalog, 8076, 0 Lakin, N. M., Pachkov, M., Tulej, M., et al. 2000, J. Chem. Phys., 113, 9586 Le Teuff, Y. H., Millar, T. J., & Markwick, A. J. 2000, A&AS, 146, 157 Liszt, H. & Lucas, R. 2001, A&A, 370, 576 Lucas, R. & Liszt, H. 1996a, A&A, 307, 237 Lucas, R. & Liszt, H. S. 1996b, in IAU Symp. 178: Molecules in Astrophysics: Probes & Processes, 421 Lucas, R. & Liszt, H. S. 2000, A&A, 358, 1069 Lykke, K. R., Neumark, D. M., Andersen, T., Trapa, V. J., & Lineberger, W. C. 1987, J. Chem. Phys., 87, 6842 Maier, J. P., Lakin, N. M., Walker, G. A. H., & Bohlender, D. A. 2001, ApJ, 553, 267 Massey, P. & Thompson, A. B. 1991, AJ, 101, 1408 McCall, B. J., Geballe, T. R., Hinkle, K. H., & Oka, T. 1998, Science, 279, 1910 McCall, B. J., Hinkle, K. H., Geballe, T. R., et al. 2002a, ApJ, 567, 391 McCall, B. J., Oka, T., Thorburn, J., Hobbs, L. M., & York, D. G. 2002b, ApJ, 567, L145 McCall, B. J., Thorburn, J., Hobbs, L. M., Oka, T., & York, D. G. 2001, ApJ, 559, L49 McCarthy, M. C., Gottlieb, C. A., Gupta, H., & Thaddeus, P. 2006, ApJ, 652, L141 Millar, T. J., Herbst, E., & Bettens, R. P. A. 2000, MNRAS, 316, 195 Millen, D. J., Topping, G., & Lide, D. R. J. 1962, J. Molo. Spec., 8, 153 Moran, S., Ellis, B. J., Defrees, D. J., Mclean, A. D., & Ellison, G. B. 1987, J. Am. Chem. Soc, 109, 5996 Morton, D. C. 2003, ApJS, 149, 205 Ozeki, H., Hirao, T., Saito, S., & Yamamoto, S. 2004, ApJ, 617, 680 Pan, K., Federman, S. R., Sheffer, Y., & Andersson, B.-G. 2005, ApJ, 633, 986 Perryman, M. A. C. & ESA. 1997, The HIPPARCOS and TYCHO cata- logues. Astrometric and photometric star catalogues derived from the ESA HIPPARCOS Space Astrometry Mission (The Hipparcos and Tycho cata- logues. Astrometric and photometric star catalogues derived from the ESA Hipparcos Space Astrometry Mission, Publisher: Noordwijk, Netherlands: ESA Publications Division, 1997, Series: ESA SP Series vol no: 1200, ISBN: 9290923997 (set)) Ruffle, D. P., Bettens, R. P. A., Terzieva, R., & Herbst, E. 1999, ApJ, 523, 678 Sailer, W., Pelc, W., Limao-Vieira, P., et al. 2003, Chem. Phys. Lett., 381, 216 Sarre, P. J. 1980, J. de Chim. Phys., 77, 769 Sarre, P. J. 2000, MNRAS, 313, L14 Sarre, P. J. 2006, J. Mol. Spec., 238, 1 Sarre, P. J. & Kendall, T. R. 2000, in IAU Symp. 197: Astrochemistry: From Molecular Clouds to Planetary Systems, 343 Savage, B. D. & Sembach, K. R. 1991, ApJ, 379, 245 Scappini, F., Casu, S., Cecchi-Pestellini, C., & Olberg, M. 2002, MNRAS, 337, 495 Sommerfeld, T. 2005, Journal of Physics: Conference Series, 4, 245 Stokes, G. M. 1978, ApJS, 36, 115 Takakuwa, S., Kawaguchi, K., Mikami, H., & Saito, M. 2001, PASJ, 53, 251 Terzieva, R. & Herbst, E. 2000, International Journal of Mass Spectrometry, 201, 135 Tulej, M., Kirkwood, D. A., Pachkov, M., & Maier, J. P. 1998, ApJ, 506, L69 Turner, B. E., Friberg, P., Irvine, W. M., Saito, S., & Yamamoto, S. 1990, ApJ, 355, 546 Tyler, J., Sheridan, J., & Costain, C. C. 1972, J. Mol. Spectrosc., 43, 248 van Dishoeck, E. F. & Black, J. H. 1986, ApJS, 62, 109 Wegner, W. 1994, MNRAS, 270, 229 Wegner, W. 2006, MNRAS, 371, 185 Welty, D. 1998, LNP Vol. 506: IAU Colloq. 166: The Local Bubble and Beyond, 506, 151 Welty, D. E. & Hobbs, L. M. 2001, ApJS, 133, 345 Welty, D. E., Hobbs, L. M., & Morton, D. C. 2003, ApJS, 147, 61 Welty, D. E., Morton, D. C., & Hobbs, L. M. 1996, ApJS, 106, 533
astro-ph/9701208
1
9701
1997-01-27T16:56:25
Early Objects in the Cosmic String Theory with Hot Dark Matter
[ "astro-ph" ]
We study the accretion of hot dark matter onto moving cosmic string loops, using an adaptation of the Zeldovich approximation to HDM. We show that a large number of nonlinear objects of mass greater than $10^{12}M_\odot$, which could be the hosts of high redshift quasars, are formed by a redshift of $z=4$.
astro-ph
astro-ph
ASTROPHYSICS REPORTS (Special Issue) Publ. Beijing Astronomical Observatory No. 2, Page 00–00 January 1997 EARLY OBJECTS IN THE COSMIC STRING THEORY WITH HOT DARK MATTER Richhild Moessner1 and Robert Brandenberger2 1Max Planck Institut fur Astrophysik, 85740 Garching, Germany 2 Brown University, Department of Physics, Providence, RI 02912, USA ABSTRACT We study the accretion of hot dark matter onto moving cosmic string loops, using an adaptation of the Zeldovich approximation to HDM. We show that a large number of nonlinear objects of mass greater than 1012M⊙, which could be the hosts of high redshift quasars, are formed by a redshift of z = 4. Key words cosmic strings, cosmology: theory, large-scale structure of Universe 1 Introduction Topological defect theories provide an alternative to inflation for explaining the origin of the primordial fluctuations which have grown into present-day structures through gravitational instability. Cosmic strings are one-dimensional topological defects possibly formed in a phase transition in the early Universe. The observed presence of massive nonlinear structures at high redshifts of 3 to 4 has provided stringent constraints on models of structure formation containing hot dark matter (HDM), a candidate for which is a massive neutrino. This is due to the large thermal velocities of the HDM particles, which prevent the growth of perturbations on scales smaller than the free- streaming length. For inflationary models with adiabatic density fluctuations, recent data on the abundance of damped Lyman alpha absorption systems (DLAS)[1] and on the quasar abundance[2] has restricted the fraction of HDM to less than about 30% of the total dark matter present[3]. The scenario of structure formation with cosmic strings is, however, viable even if the dark matter in the universe is hot, because cosmic strings, which provide the seeds for the density perturbations, survive the neutrino free-streaming[4, 5]. In this article we present work on the constraints imposed on the cosmic string theory by the abundance of high redshift quasars[6]. Searches for high redshift quasars have been going on for some time[2]. The quasar luminos- ity function is observed to rise sharply as a function of redshift z until z ≃ 2.5. According to recent results from the Palomar grism survey by Schmidt et al. (1995)[2], it peaks in the redshift interval z ǫ [1.7, 2.7] and declines at higher redshifts. Quasars (QSO) are extremely luminous, and it is generally assumed that they are powered by accretion onto black holes. It is possible to estimate the mass of the host galaxy of the quasar as a function of its luminosity, assuming that the quasar luminosity corresponds to the Eddington luminosity of the black hole. For a quasar of absolute blue magnitude MB = −26 and lifetime of tQ = 108yrs, the host galaxy mass can be estimated as[7] MG = c11012M⊙ , (1) Early Objects from Cosmic Strings 1 where c1 is a constant which contains the uncertainties in relating blue magnitude to bolometric magnitude of quasars, in the baryon fraction of the Universe and in the fraction of baryons in the host galaxy able to form the compact central object (taken to be 10−2). The best estimate for c1 is about 1. Models of structure formation have to pass the test of producing enough early objects of sufficiently large mass to host the observed quasars. Besides the quasars themselves, absorption lines in their spectra due to intervening matter can be used to quantify the amount of matter in nonlinear structures at high redshifts. Based on the number density of absorption lines per frequency interval and on the column density calculated from individual absorption lines, the fraction of Ω in bound neutral gas (denoted by Ωg) can be estimated. For high-column density systems, the damped Ly-α systems (DLAS), recent observational results are that Ωg(z) > 10−3 (2) for z ǫ [1,3]. In the most recent results by Storrie-Lombardi et al. (1996)[1] there is evidence for a flattening of Ωg at z ∼ 2 and a possible turnover at z ∼ 3 . The corresponding value for Ω in bound matter is larger by a factor of f −1 , where fb is the fraction of bound matter which is baryonic, which is about 10% in a flat universe. b In the next section, we give a brief review of the cosmic string scenario of structure formation. Then we study the accretion of hot dark matter onto cosmic string loops, which seed large amplitude local density contrasts already at early times. We use the results to compute the number density of high redshift objects as a function of a parameter ν which determines the number density of loops in the scaling solution (see Section 2). We demonstrate that for realistic values of ν, the number of massive nonlinear objects at redshifts ≤ 4 satisfies the recent observational constraint of quasar abundances (see Figure 1). We also comment on the implications of (2). We consider a spatially flat Universe containing HDM and baryons. Units where c = 1, a Hubble constant of H = 50 h50kms−1Mpc−1 and a redshift at equal matter and radiation of zeq = 5750 Ωh2 50 are used. 2 Cosmic Strings and Structure Formation Cosmic strings[8] are one-dimensional topological defects which might have been formed in a phase transition in the very early Universe. They are produced for the same reasons as ordinary defects in condensed matter systems, such as vortex lines in superfluid helium, but their subsequent dynamics is governed by relativistic equations. In the simplest model, the symmetry breaking occuring in the phase transition is achieved via a two-component scalar Higgs field φ = (φ1, φ2) with symmetry-breaking potential V (φ) = 1 4 λ(φ2 − η2)2, where η is the energy scale of the phase transition, and λ is the coupling constant. Topological considerations require cosmic strings to have no ends, they are either formed as infinitely long strings or as closed loops. These strings are very thin lines of trapped energy, the energy being the false vacuum energy V (φ = 0) which is trapped inside the defects after the rest of the Universe has evolved to the true ground state < φ > = η with V (φ) = 0. The mass per unit length of the strings is given by the symmetry breaking scale η , µ ∼ η2 , and it is the only free parameter of the model (in practice uncertainties due to the complicated evolution of cosmic strings after the phase transition introduce extra phenomenological parameters, which are not fundamental, however). The energy in the defects acts gravitationally on surrounding matter and radiation, and thus provides the origin for the density perturbations, which evolve into today's large-scale structure. Most important for structure formation are the infinite strings. They are straight over distances increasing with time of one horizon distance H −1(t) ∼ t, and form an approximate 2 R. Moessner & R. Brandenberger random walk on distances larger than that. In this way they can cause structures to be formed on the large scales observed today. The model makes predictions consistent with large-scale structure observations[9, 10] and CMB anisotropy measurements[11]. An intriguing fact is that the normalization of the free parameter µ of the model to these observations, which gives Gµ ≈ 10−6, requires the phase transition giving rise to the defects to take place at a scale of η ∼ 1016GeV . This is just the scale of grand unification where the three coupling constants of the strong, weak and electromagnetic interactions meet, suggesting the appearance of new physics. The network of cosmic strings is formed in a phase transition in the early universe about 10−35 sec after the Big Bang, long before the times relevant for structure formation. Therefore the string network has to be evolved over many orders of magnitude in time until teq (and subse- quently). This is very complicated numerically, and would be impossible to handle analytically, if it were not for the fact that the network of strings quickly evolves towards a scaling solution[8] where the energy density in long strings remains a constant fraction of the total background energy density. This is achieved by intercommutations and self-intersections of the strings lead- ing to the production of small loops, which then decay by emitting gravitational radiation. In this way, some of the energy input into the string network coming from the stretching of the strings due to the expansion of the universe is transferred to the background. According to the scaling solution, the string network looks the same at all times when all distances are scaled by the Hubble radius H −1(t), and this allows the extrapolation over such long time intervals. The scaling solution can be pictured as having a fixed number M of long strings per Hubble volume at any given time, and a distribution of small loops. Some of the parameters characterizing the string network besides the fundamental one µ are the long string velocities vs and initial loop velocities vi, the average number M of strings per Hubble volume, the amount of small-scale structure on the long strings quantified by (µ − T ), where T is the tension of the string, and the constants α, β and ν appearing in the next three equations. Small-scale structure is produced mainly when string segments self-interect and split off loops, a process which is characterized by the loop production rate and the size of the produced loops. According to the scaling solution, loops are formed with a radius which is a constant fraction of the horizon, and length Rf (t) = αt , l = βR . (3) (4) The number of loops per unit physical volume present at time t with lengths in the interval between l and l + dl is n(l, t)dt = νl−2t−2dt . (5) Since loops decay by emitting gravitational radiation, there is a lower cutoff value of l for the distribution (5) given by lmin ∼ Gµt, G being Newton's constant. Below lmin, the distribution n(l, t) becomes constant. Numerical simulations[12] indicate that α ≤ 10−2, β ≃ 10, and M ∼ 10. From these values it follows that – unless ν is extremely large – most of the mass of the string network resides in long strings (where long strings are defined operationally as strings which are not loops with radius smaller than the Hubble radius). Long strings accrete matter in the form of wakes behind them as they move through space[13, 9, 10]. Spacetime around a long straight cosmic string can be pictured as locally flat, but with a deficit angle[14] of 8πGµ . Therefore a string moving relativistically with ve- locity vs imparts velocity perturbations to surrounding matter towards the plane swept out by Early Objects from Cosmic Strings the string, of magnitude u = 4πGµγsvs 3 (6) creating overdensities in the form of a wake behind the string. If small-scale structure is present on the string, there is in addition a Newtonian force towards the string, proportional to G(µ−T ), the strings move more slowly and accrete matter rather in the form of filaments than wakes. For HDM, the first nonlinearities about wakes resulting from strings without any small- scale structure form only at late times, at a redshift of[10] about 1 for Gµ ≃ 10−6, which is obtained when normalizing the model to COBE. Before this redshift, no nonlinearities form as a consequence of accretion onto a single uniform wake. Thus, in the cosmic string and hot dark matter theory, a different mechanism is required in order to explain the origin of high redshift objects. Possible mechanisms related to wakes are early structure formation at the crossing sites of different wakes[15], small-scale structure of the strings giving rise to wakes[16, 17], and inhomogeneities inside of wakes[18]. In this article, however, we will explore a different mechanism, namely the accretion of hot dark matter onto loops. In earlier work[5], the accretion of hot dark matter onto static cosmic string loops was studied. It was found that in spite of free streaming, the nonlinear structure seeded by a point mass grows from inside out, and that the first nonlinearities form early on (accretion onto string filaments proceeds similarly[17]). Since loop accretion leads to nonlinear structures at high redshift, we will now investigate this mechanism in detail to see whether enough high redshift massive objects to statisfy the QSO constraints and (2) form. 3 Nonlinear mass accreted by a single static loop We will use the modified Zeldovich approximation[19, 10] to study the accretion of hot dark matter onto moving string loops. The Zeldovich approximation is a first order Lagrangian perturbation theory technique in which the time evolution of the comoving displacement ψ of a dark matter particle from the location of the seed perturbation is studied. The physical distance of a dark matter particle from the center of the cosmic string loop is written as h(q, t) = a(t)[q − ψ(q, t)] . For a point-like seed mass of magnitude m (the string loop in our case) located at the comoving position q′ = 0, the equation for ψ is (cid:18) ∂2 ∂t2 + 2 a a ∂ ∂t − 4πGρ(t)(cid:19) ψ(q, t) = Gm a3q2 . (7) This equation describes how as a consequence of the seed mass, the motion of the dark matter particles away from the seed (driven by the expansion of the Universe) is gradually slowed down. If the seed perturbation is created at time ti and the dark matter is cold, then the appropriate initial conditions for ψ are ψ(q, ti) = ψ(q, ti) = 0 . (8) As formulated above, the Zel'dovich approximation only works for cold dark matter, particles with negligible thermal velocities. For HDM, free-streaming prevents the growth of perturba- tions on scales q below the free-streaming length λJ (t), which is the mean comoving distance traveled by neutrinos in one expansion time λJ (t) = v(t)z(t)t , 4 R. Moessner & R. Brandenberger where v(t) ∼ z(t) is the hot dark matter velocity. The free- streaming length decreases with time (in comoving coordinates) as t−1/3, so that a scale q which is initially affected by free- streaming, q < λJ (ti) becomes equal to λJ at some later time ts(q) , and then remains above it and unaffected by free-streaming at later times. A correct treatment of the effect of free-streaming requires the solution of the collisionless Boltzmann equation for the neutrino phase-space density[5]. But a simple modification of the Zeldovich approximation for CDM gives the same results as the full treatment[10]. For scales q < λJ (ti), Eq. 7 is only integrated from the time ts(q) > ti onwards instead of from ti , i.e. growth only starts at ts(q) , and the initial conditions are modified to ψ(q, ts(q)) = 0 and ψ(q, ts(q)) = 0 . For scales where q > λJ (ti) initially the CDM formalism applies. We can now define the mass that has gone nonlinear about a seed perturbation as the rest mass inside of the shell which is "turning around", i.e. for which In the case of CDM, this yields a nonlinear mass of h(q, t) = 0 . MCDM (t) = m(cid:18) t ti(cid:19)2/3 2 5 , and for loops where the accretion is affected by free-streaming, MHDM (t) = 8 125 m3 M 2 eq (cid:18) t teq(cid:19)2 (9) (10) with Meq = 2v3 eqteq/(9G) , veq is the HDM-particle velocity at teq. From Eq. 9 and 10 we can determine, for any time t, a mass M ′(t) such that for all masses M > M ′(t) accreted by any loop so far, accretion has not been affected by free-streaming, and Eq. 9 is applicable. For M < M ′(t), expression 10 is valid. 4 Number density of quasar host galaxies We can now compute the number density nG(> MG, t) in nonlinear objects heavier than MG at high redshift z(t) in the cosmic string and hot dark matter model (in an analogous way the fraction Ωnl of the critical density in such objects can be calculated). By considering only the accretion onto string loops we will be underestimating these quantities. It can be shown[20] that in an HDM model string loops accrete matter independently, at least before the large-scale structure in wakes turns nonlinear at a redshift of about 1, i.e., later than the times of interest in this paper. Hence, the number density n(l, t) of loops of length l given by (5) can be combined with the mass M (l) accreted by an individual loop to give the mass function n(M, t). Here, n(M, t)dM is the number density of objects with mass in the interval between M and M + dM at time t. This in turn determines the comoving density in objects of mass M > MG. Since the functional form of M (m) and hence M (l) changes at M = M ′, the functional forms of the mass function n(M ) will be different above and below M ′. Thus M ′(t) M2(t) nG(> MG, t) = z−3(t)  Z MG dM nH (M ) + Z M ′(t) dM nC (M )  . (11) Early Objects from Cosmic Strings 5 Similarly, the fraction of the critical density in objects of mass greater than M1, Ωnl(t) is Ωnl(t) =   M ′(t) M2(t) Z M1 dM nH(M )M + Z M ′(t) dM nC (M )M  6πGt2 . (12) Note the upper mass cutoff M2(t) introduced. It is imposed by our approximation of loops as point masses. For masses greater than M2(t), this approximation breaks down, the responsible loops would have too large a radius to be able to accrete matter effectively. M2(t) can be estimated by demanding that the mass M (t) accreted onto a loop exceed the mass in a sphere of radius equal to the loop radius at ti . The integral in Eq. 11 is dominated at M ′(t) , and can be approximated as nG(> MG, t) ∼ nC (M ′, t)M ′z−3(t) ≃ 1 5 ν(αβ)2 µ3 M ′(t)3 z−3(t) . (13) In Figure 1, the comoving number density of objects massive enough to be able to host quasars, nG(> MG, t), is plotted for Gµ = 10−6, veq = 0.1, and the values of the parameters α = 10−2 and β = 10. The upper curve is for static loops, the lower one for initial loop velocities of vi = 0.25. The effect of loop motion is to decrease the density of these massive objects, as discussed in the next paragraph. We have to compare our results for nG(> MG, t) with the number density of host galaxies, nobs G , inferred from the observed quasar abundance nQ. Since we assumed a finite quasar lifetime tQ, the number of host galaxies nobs G is larger than nQ by a factor of tH/tQ, where tH is the Hubble time at time t, nobs G = tH tQ nQ . (14) This is also plotted in Fig. 1 . The formalism presented above can be adapted to moving instead of static loops[6, 21]. Loop motion changes the geometry of the accreted object, it becomes more elongated and the transverse scale going nonlinear becomes smaller than for loops at rest. Because of this, free-streaming of the neutrinos can hinder the growth of perturbations for a longer time. The smallest mass accreted at time t which has been unaffected by free-streaming, M ′(t), is increased compared to the case of static loops, thereby decreasing nG(t) according to Eq. 13. The result for the fraction of nonlinear mass accreted by moving loops is Ωnl(t) ∼ 10−2να−2(Gµ)2 6v−1 i z(t)−2h4 50 , (15) where (Gµ)6 is defined as (Gµ)/10−6. In Fig. 1 we have plotted the curves up to a redshift of 5. There is another issue to consider, however. Our expresions for nG are only valid as long as M ′ < M2(t), due to our approximation of loops as point sources. For loops with an initial velocity of vi = 0.25, this condition is satisfied only up to a largest redshift of zmax = 6 5 βGµzeqv−1 i v−1 eq , (16) which is equal to zmax = 3h2 50 (17) for vi = 0.25 and veq = 0.1 . Beyond that redshift, the values of nG(t) and Ωnl are suppressed beyond the results plotted in Fig. 1 and Eq. 15 since only the tail of the loop ensemble with velocities smaller than the mean velocity vi = 0.25 manage to accrete a substantial amount of mass. It is interesting to note that, as mentioned in the introduction, Storrie-Lombardi et al. (1996)[1] find a flattening of Ωg at z ∼ 2 and a possible turnover at z ∼ 3. 6 R. Moessner & R. Brandenberger 5 Discussion We have studied the accretion of hot dark matter onto moving cosmic string loops and made use of the results to study early structure formation in the cosmic string plus HDM model. The loop accretion mechanism is able to generate nonlinear objects which could serve as the hosts of high redshift quasars much earlier than the time cosmic string wakes start becoming nonlinear (which for Gµ = 10−6 and vs = 1/2 occurs at a redshift of about 1). The fraction Ωnl(z) of the total mass accreted into nonlinear objects by string loops unfor- tunately depends very sensitively on α and ν. On the other hand, this is not surprising since the power of the loop accretion mechanism depends on the number and initial sizes of the loops, and the scaling relation Ωnl ∼ να is what should be expected from physical considerations. For the values ν = 1 and α−2 = 1 which are indicated by recent cosmic string evolution simulations[12], we conclude that the loop accretion mechanism produces enough large mass protogalaxies to explain the observed abundance of z ≤ 4 quasars (see Figure 1). Note that the amplitude of the predicted protogalaxy density curves depends sensitively on the parameters of the cosmic string scaling solution which are still poorly determined. Hence, the important result is that there are parameters for which the theory predicts a sufficient number of protogalaxies. Since not all protogalaxies will actually host quasars, and since the string parameters are still uncertain, it would be wrong to demand that the amplitude of the nG curve agree with that of the observed nQ; rather, it should lie above the nQ curve. It is more difficult to make definite conclusions regarding the abundance of damped Lyman alpha absorption systems. In the form of Eq. 15, the condition for the cosmic string loop accretion mechanism to be able to explain the data is also satisfied. However, Eq. 15 refers to the value of Ω in baryonic matter. The corresponding constraint on the total matter collapsed in structures associated with damped Lyman alpha systems is ΩDL(z < 3) > f −1 b 10−3 where fb is the local fraction of the mass in baryons. From Eq. 5 it follows that the above constraint is only marginally satisfied, and this only if the local baryon fraction fb exceeds the average value for the whole Universe of about fb = 0.1. But in the cosmic string model with HDM we might expect fb in nonlinear objects to be enhanced over the average fb because after teq baryons are able to cluster during the time that the HDM is prevented from accreting by the free streaming. Thus, cosmic strings may be able to restore agreement with ( 15) in a natural way. More calculations are required to resolve this issue. Here we have only reported on the mechanism of forming early nonlinear objects through accretion onto string loops. Another mechanism is through small-scale structure on the long strings leading to the formation of filaments rather than wakes, which has recently been investigated[22]. It was found that this could be the most effective mechanism, and for the maximal possible amount of small-scale structure, Ωnl ∼ 1 can be reached already at a redshift of 5. It is clearly important to determine the amount of small-scale structure present on strings. 6 Acknowledgements R.M. wants to thank Gerhard Borner and Deng Zugan for the invitation to attend this workshop. We are grateful to Martin Hahnelt and Houjun Mo for useful discussions. References Early Objects from Cosmic Strings 7 [1] K. Lanzetta et al., Ap. J. (Suppl.) 77, 1 (1991); K. Lanzetta, D. Turnshek and A. Wolfe, Ap. J. (Suppl.) 84, 1 (1993); L. Storrie-Lombardi, R. McMahon, M. Irwin and C. Hazard, 'High redshift Lyman limit and damped Lyman alpha absorbers', astro-ph/9503089, to appear in 'ESO Workshop on QSO Absorption Lines' (1995); L. Storrie-Lombardi, R. McMahon, M. Irwin,MNRAS 283, L79 (1996). [2] S. Warren, P. Hewett and P. Osmer, Ap. J. (Suppl.) 76, 23 (1991); M. Irwin, J. McMahon and S. Hazard, in 'The space distribution of quasars', ed. D. Crampton (ASP, San Francisco, 1991), p. 117; M. Schmidt, D. Schneider and J. Gunn, ibid., p. 109; B. Boyle et al., ibid., p. 191; M. Schmidt, D. Schneider and J. Gunn, A. J. 110, 68 (1995). [3] C.P.. Ma, E. Bertschinger, Ap. J. 434, L25 (1994); H.J. Mo, J. Miralda-Escude, Ap. J. 430, L25 (1994); A. Klypin, S. Borgani, J. Holtzman and J. Primack, Ap. J. 444, 1 (1995). [4] A. Vilenkin and Q. Shafi, Phys. Rev. Lett. 51, 1716 (1983). [5] R. Brandenberger, N. Kaiser, D. Schramm and N. Turok, Phys. Rev. Lett. 59, 2371 (1987); R. Brandenberger, N. Kaiser and N. Turok, Phys. Rev. D36, 2242 (1987). [6] R. Moessner and R. Brandenberger, MNRAS 280, 797 (1996), astro-ph/9510141 . [7] K. Subramanian and T. Padmanabhan, 'Constraints on the models for structure formation from the abundance of damped Lyman alpha systems', IUCAA preprint, astro-ph/9402006 (1994). [8] A. Vilenkin and E.P.S. Shellard, 'Cosmic strings and other topological defects' (Cambridge Univ. Press, Cambridge, 1994); M. Hindmarsh and T. Kibble, Rep. Prog. Phys. 58, 477 (1995).; R. Brandenberger, Int. J. Mod. Phys. A9, 2117 (1994). [9] T. Vachaspati, Phys. Rev. Lett. 57, 1655 (1986). [10] L. Perivolaropoulos, R. Brandenberger and A. Stebbins, Phys. Rev. D41, 1764 (1990); R. Brandenberger, L. Perivolaropoulos and A. Stebbins, Int. J. Mod. Phys. A5, 1633 (1990). [11] D. Bennett, F. Bouchet and A. Stebbins, Ap. J. (Lett.) 399, L5 (1992); L. Perivolaropoulos, Phys. Lett. 298B, 305 (1993). [12] D. Bennett and F. Bouchet, Phys. Rev. Lett. 60, 257 (1988); B. Allen and E. P. S. Shellard, Phys. Rev. Lett 64, 119 (1990); A. Albrecht and N. Turok, Phys. Rev. D40, 973 (1989). [13] J. Silk and A. Vilenkin, Phys. Rev. Lett 53, 1700 (1984). [14] A. Vilenkin, Phys. Rev. D23, 852 (1981). [15] T. Hara and S. Miyoshi, Prog. Theor. Phys. 81, 1187 (1989); T. Hara, S. Morioka and S. Miyoshi, Prog. Theor. Phys. 84, 867 (1990); T. Hara et al., Ap. J. 428, 51 (1994). [16] D. Vollick, Phys. Rev. D 45, 1884 (1992); T. Vachaspati and A. Vilenkin, Phys. Rev Lett. 67, 1057 (1991). [17] A. Aguirre and R. Brandenberger, astro-ph/9505031, Int. J. Mod. Phys. D4, 711 (1995). [18] M. Rees, Mon. Not. R. Astr. Soc. 222, 27P (1986). 8 R. Moessner & R. Brandenberger [19] Ya.B. Zel'dovich, Astron. Astrophys. 5, 84 (1970). [20] R. Brandenberger and E.P.S. Shellard, Phys. Rev. D40, 2542 (1989). [21] E. Bertschinger, Ap. J. 316, 489 (1987). [22] V. Zanchin, J.A.S. Lima, R. Brandenberger, Phys. Rev. D54, 7129 (1996), astro-ph/9607062 . Early Objects from Cosmic Strings 9 Figure 1: Comparison of the number density of host galaxies nobs G ('*' marks) inferred from the observed number density of quasars brigher than MB = −26 from the Schmidt et al. (1991,1995) survey('x' marks) with the number density nG of protogalaxies of mass greater than 1012M⊙ predicted in the cosmic string theory with HDM, for the parameters discussed in the text, and h50 = 1. The horizontal axis is the redshift.
0711.1354
1
0711
2007-11-08T21:00:36
A Population of Faint Extended Line Emitters and the Host Galaxies of Optically Thick QSO Absorption Systems
[ "astro-ph" ]
We have conducted a long slit search for low surface brightness Lyman-alpha emitters at redshift 2.67 < z < 3.75. A 92 hour long exposure with VLT/FORS2 down to a 1-sigma surface brightness detection limit of 8x10^-20 erg/cm2/s/sqarcsec yielded a sample of 27 single line emitters with fluxes of a few times 10^-18 erg/s/cm2. We present arguments that most objects are indeed Lyman-alpha. The large comoving number density, the large covering factor, dN/dz ~ 0.2-1, and the often extended Lyman-alpha emission suggest that the emitters be identified with the elusive host population of damped Lyman-alpha systems (DLAS) and high column density Lyman limit systems. A small inferred star formation rate, perhaps supplanted by cooling radiation, appears to energetically dominate the Lyman-alpha emission, and is consistent with the low metallicity, low dust content, and theoretically inferred low masses of DLAS, and with the relative lack of success of earlier searches for their optical counterparts. (abridged)
astro-ph
astro-ph
A Population of Faint Extended Line Emitters and the Host Galaxies of Optically Thick QSO Absorption Systems1 Observatories of the Carnegie Institution of Washington, 813 Santa Barbara Street, Michael Rauch Pasadena, CA 91101 [email protected] Martin Haehnelt Institute of Astronomy, Madingley Road, Cambridge CB3 0HA, UK [email protected] Andrew Bunker School of Physics, Stocker Road, Exeter EX4 4QL, UK [email protected] George Becker Observatories of the Carnegie Institution of Washington, 813 Santa Barbara Street, Pasadena, CA 91101, USA [email protected] Francine Marleau Spitzer Science Center, Caltech, Mail Stop 220-6, 1200 East California Blvd, Pasadena, CA 91125, USA [email protected] James Graham 601 Campbell Hall, University of California, Berkeley CA 94720-3411 ,USA [email protected] Stefano Cristiani Osservatorio Astronomico di Trieste, INAF, Via Tiepolo 11 34143 Trieste, Italy – 2 – [email protected] Matt J. Jarvis Centre for Astrophysics, Science & Technology Research Institute, University of Hertfordshire, Hatfield AL10 9AB, UK [email protected] Cedric Lacey Institute for Computational Cosmology, Department of Physics, Durham University, South Road, Durham DH1 3LE, UK [email protected] Simon Morris Department of Physics, Durham University, South Road, Durham DH1 3LE, UK [email protected] Celine Peroux Observatoire Astronomique de Marseille-Provence, 2, Place Le Verrier,13248 Marseille Cedex 04, France [email protected] Huub Rottgering Leiden Observatory, NL-2300 RA Leiden, The Netherlands [email protected] Tom Theuns Department of Physics, Durham University, South Road, Durham DH1 3LE, UK [email protected] – 3 – ABSTRACT We have conducted a long slit search for low surface brightness Lyman α emit- ters at redshift 2.67 < z < 3.75. A 92 hour long exposure with the ESO VLT FORS2 instrument down to a 1σ surface brightness detection limit of 8 × 10−20 erg cm−2 s−1 ⊓⊔′′−1 per ⊓⊔′′ aperture, yielded a sample of 27 single line emitters with fluxes of a few ×10−18 erg s−1cm−2. We present arguments that most ob- jects are indeed Lyα. The large comoving number density 3 × 10−2 h3 70 Mpc−3, the large covering factor dN/dz ∼ 0.2 − 1, and the often extended Lyα emis- sion suggest that the emitters be identified with the elusive host population of damped Lyα systems (DLAS) and high column density Lyman limit systems (LLS). A small inferred star formation rate, perhaps supplanted by cooling ra- diation, appears to energetically dominate the Lyα emission, and is consistent with the low metallicity, low dust content, and theoretically inferred low masses of DLAS, and with the relative lack of success of earlier searches for their optical counterparts. Some of the line profiles show evidence for radiative transfer in galactic outflows. Stacking surface brightness profiles we find emission out to at least 4′′. The centrally concentrated emission of most objects appears to light up the outskirts of the emitters (where LLS arise) down to a column density where the conversion from UV to Lyα photon becomes inefficient. DLAS, high column density LLS, and the emitter population discovered in this survey appear to be different observational manifestations of the same low-mass, protogalactic building blocks of present day L∗ galaxies. Subject headings: galaxies: formation - intergalactic medium - diffuse radia- tion - quasars: individual(DMS2139.0-0405) 1. Introduction Perhaps surprisingly, modern astronomy has found ways to study the material structures of the high redshift universe over their entire vast range of densities and sizes, from the 1Based partly on observations made with ESO Telescopes at the Paranal Observatories under Program ID LP173.A-0440, and partly on observations obtained at the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agreement with the NSF on behalf of the Gemini partnership: the National Science Foundation (United States), the Science and Technology Facilities Council (United Kingdom), the National Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), CNPq (Brazil) and CONICET (Argentina). – 4 – underdense large voids, via bright starforming galaxies to the most luminous QSOs. The dark stages of this sequence, with densities from the universal mean to virial density, too low to produce detectable radiation, have been studied mainly with QSO absorption lines. This approach has taught us where to find most of the baryons (in the ionized Lyman α forest gas; e.g., Rauch et al 1997a) and most of the neutral gas (in damped Lyman α systems; e.g., Wolfe et al 1995) at high redshift. From the bright end of the cosmic matter distribution, serious inroads into the population of high redshift galaxies have been made by color-selecting stellar continuum emitters, exploiting the Lyman limit continuum decrement (e.g., Steidel & Hamilton 1993), or by searching for Lyα line emission induced by star-formation (e.g., Cowie & Hu 1998). The advent of space-based, broad-band galaxy surveys, in particular the Hubble Ultra Deep Field (Bunker et al 2004; Beckwith et al 2006; Bouwens et al 2007) has begun filling in the gap between the essentially dark, lower mass range of sub-galactic or barely star-forming proto-galactic objects probed by absorption lines, and bright high redshift galaxies with large star-formation rates of tens of M⊙yr−1 (e.g., Erb et al 2006). The boundary between dark and bright is delineated by an important astrophysical transition, the one from ionized gas to neutral, self-shielded gas, on a galactic mass scale. The underlying objects in which this transition has happened, known from absorption studies as Lyman limit systems (LLS), or, at higher column densities, damped Lyα systems (DLAS), have been rather elusive. We know quite a bit about their chemistry and ionization state but little in terms of stellar contents, size, kinematics or mass. Theoretical studies over the past two decades have suggested that a survey of LLS and DLAS for HI Lyα line emission of sufficient depth may uncover several distinct astrophysical sources of line emission that have the potential to shed considerable light (literally) on the distribution of neutral hydrogen, and thus the bedrock of galaxy formation. If we acknowl- edge that high redshift DLAS must have something to do with stars (they contain the main reservoir of neutral gas, and have a median metallicity of [Z/H] = -1.5, more than an order of magnitude higher than the coeval abundances in the intergalactic medium) then we may expect the potentially strongest signal to be star formation induced Lyα and/or a stellar continuum. Attempts at identifying individual LLS or DLAS with galaxy counterparts have often been frustrated by the difficulty of detecting an extremely faint object (the DLAS host) next to an extremely bright object (a QSO). At low redshift (z < 1) where we can hope to learn most about the underlying galaxy population, observations have shown that DLA host galaxies represent a range of galaxy types (e.g., LeBrun et al 1997, Chen, Kennicutt & Rauch 2005) dominated by faint objects (e.g., Chun et al 2006) with low star formation rates (Wild, Hewett & Pettini 2007). Searches at high redshift (e.g., Warren et al 2001, Fynbo et al 2003; Kulkarni et al 2000, 2001; Christensen et al 2007) have so far produced only a – 5 – handful of confirmed detections of the underlying galaxies (Weatherley et al 2005). Such efforts indicated that DLAS hosts at high redshift are generally drawn from the very faint end (Fynbo, Møller, & Warren 1999; Bunker et al 1999) of the general galaxy population at high redshift, intersected by the QSO line of sight at small impact parameters (Møller et al 2002). Wolfe & Chen (2007) have recently performed a search for spatially extended continuum emission down to very faint levels using the Hubble Ultra Deep Field, and were able to place stringent upper limits on extended star formation in DLAS. These results agree well with theoretical CDM-based models of galaxy formation (Kauff- mann 1996) that envisage DLAS hosts as numerous small, faint, low mass, merging proto- galactic clumps (Haehnelt, Steinmetz & Rauch 1998, 2000; Ledoux et al 1998; Johansson & Efstathiou 2006; Nagamine et al 2007), rather than the large hypothetical disks (Prochaska & Wolfe 1997) once popular. Cooling radiation from collapsing galaxies (e.g., Haiman & Rees 2001; Haiman, Spaans & Quataert 2000; Fardal et al 2001, Furlanetto et al 2005) is a second process able to produce line radiation at fluxes competitive with those from star formation, at least in the halos of relatively massive galaxies (Dijkstra et al 2006a,b). The peculiar spatial distribution of the radiation and certain asymmetries of the spectral line profile, together with the local absence of a stellar continuum, could conceivably help to distinguish this source of radiation from star formation. The few candidates for such galactic cooling flows reported so far (Keel et al 1999; Steidel et al 2000; Francis et al 2001; Bower et al 2004; Dey et al 2005; Matsuda et al 2006; Nilsson et al 2006; Smith & Jarvis 2007), have been atypical for the galaxy population as a whole. Perhaps the most curious of these emission processes is Lyα fluorescence, where HI ion- izing photons impinging on optical thick gas are absorbed and converted with high efficiency to Lyα line radiation, albeit at a very low surface brightness level < 10−19erg cm−2s−1⊓⊔′′−1; Hogan & Weymann 1987; Binette et al 1993; Gould & Weinberg 1996; Cantalupo, Lily & Porciani 2005). Lyman limit patches should light up the whole gaseous cosmic web to yield a relatively uniform, very faint "glow", perhaps locally enhanced by the proximity of a QSO (e.g., Cantalupo et al 2005, 2007). A number of increasingly deep searches for this effect at high redshift have been performed (Lowenthal et al 1990; Bunker et al 1998, 1999), in pursuit of a repeatedly shrinking theoretically predicted signal. A detection, which at the same time would be a measurement of the general UV background, has so far eluded us. Most recently, a number of detections of the enhanced Lyα fluorescence in the proximity of QSOs have been reported (Fynbo, Møller, & Warren 1999; Bunker et al 2003; Weidinger et al 2004, 2005; Adelberger et al 2006, Cantalupo et al 2007, Hennawi 2007) but the understanding of the effect has been complicated by the large number of degrees of freedom in the properties and – 6 – behavior of the QSO (e.g., orientation, opening angle of the beam, life time). Thus there is good reason to conduct a search for extremely low light level Lyα line emission at high redshift. While most previous Lyα surveys have been performed as narrow band imaging searches it is clear that the largest contrast with the sky background and the largest redshift (=spatial) depth is obtained by (long) slit spectroscopy. Because of the small volume covered a blind search is not a viable way to find objects as rare as Lyman break galaxies, which are hard to hit with a single randomly positioned slit. In contrast, the rate of incidence of LLS with neutral hydrogen column densities exceeding (N(HI) > 1019cm−2) per unit redshift is approximately unity at redshift ∼ 3 (e.g., Peroux et al 2003), i.e., the objects essentially cover the sky, and there should be numerous hits in a single setting for a typical long slit spectrograph. We anticipate seeing about 30 patches of emission on a 2′′ wide and 7′ long slit at redshift ∼ 3 (e.g., Bunker et al 1998). Searching over a redshift range near z ∼ 3 achieves a reasonable comprise between avoiding the ravages of (1 + z)4 dimming and succumbing to the relatively poor blue performance of most low resolution spectrographs worldwide. However, in principle it is desirable to go as far to the blue as possible. The UV background as the source of the Lyα fluorescence is not expected to drop dramatically down to at least redshift 2, whereas the atmospheric background as the dominant source of noise decreases rapidly toward the blue. Prior to the current ESO VLT project the most sensitive search for Lyα fluorescence off optically thick gas, in regions dominated by the general UV background, had been performed by the longslit experiment of Bunker et al (1998, 1999). The first attempt at measuring this effect with the ESO VLT originated in a (shelved) science verification project for FORS, later submitted unsuccessfully during periods 64-66 as a large project (PI Rauch). The amount of exposure time required even with an 8m class telescope considerably exceeds typical observing time awards so the strategy was to exploit the newly available service mode observing and insert the observations during bad seeing periods, as the targets were expected to be extended, with radii of several arcseconds. The current incarnation, a combination of the two projects, re-observing the original target field of Bunker et al (1999), was awarded 120 hours as ESO Large Project (LP173.A-0440: PI Haehnelt). A parallel attempt to use the Gemini GMOS instruments on the same field chosen for the ESO project led to time awards with both Gemini telescopes (GN 2004-A-Q-91, GN-2005-B-Q-52, GS-2004-B-Q-61, GS-2005-B-Q-36; PI Bunker) and (GN-2004-B-Q-35,GN-2004-B-Q-35, GS-2004A-Q-78, GS- 2004-B-Q-8: PI Rauch). The Gemini effort resulted in 46 × 3000s exposures, of which 30 where taken with Gemini-North, 16 with Gemini-South. The Gemini data were taken at lower resolution (to take advantage of the most commonly used spectroscopic setup) but cover a larger redshift path than the ESO data. The current paper deals with the ESO dataset and we postpone a full analysis of all data to a future paper. However, the Gemini – 7 – data in a preliminary reduction have been consulted here for checking the reality of the emitters found in the ESO dataset (see below). The ESO FORS2 project resulted in (after overheads) 92 hours of on-source exposure, finally reaching a 1σ surface brightness detection threshold of 8 × 10−20 erg cm−2 s−1 ⊓⊔′′−1 (measured in a one ⊓⊔′′ aperture). This is remarkably close to our expected sensitivity (6.6× 10−20 erg cm−2 s−1 ⊓⊔′′−1 in 120 h). The experiment is described below. Section 2 details the observational setup, and the sensitivity and spectral resolution limits reached. Section 3 spells out the individual proper- ties of the emitters found, and gives estimates of the sizes, number densities, flux distribu- tions, and derived rates of incidence. Section 4 discusses the possible identification with low redshift galaxies, and Section 5 ponders the consequences if the emitters are predominantly Lyα. Alternative origins of the Lyα photons (stellar, QSO-induced, cooling radiation) are discussed in section 6, including a detailed discussion of some unusual line profiles found. Section 7 establishes a connection between the emitters and QSO DLAS, followed by a brief discussion of Lyα fluorescence in section 8 and the conclusions in section 9. An Appendix discusses the importance of slit losses. 2. Observations The QSO DMS B 2139-0405 (z=3.32, V =20.805; Hall et al 1996) was observed dur- ing 2004 - 2006 with the VLT FORS2 low resolution spectrograph and the volume-phased holographic grism 1400V. A 2′′ wide and about 453′′ long slit gave a spectral resolution of λ/∆λFWHM = 1050. The CCDs were readout in 2 × 2 binned mode, giving pixels with an extent of 0.252′′ along the slit and about 0.64 A in the dispersion direction. Thus, the spec- tral resolution FHWM is sampled by about 8 pixels. The spectrum on the detector ranges from 4457 to 5776 A, with a midpoint at 5099 A. The slit was centered on the QSO and rotated by a position angle −5.95 deg, to repeat the orientation of the earlier Keck LRIS observation by Bunker et al (1999), where the particular position angle was chosen to minimize intersecting bright foreground galaxies. A total of 110 exposures were taken between May 2004 and August 2006, giving a total integration time of 92 hours. The exposures were divided roughly evenly into three dither positions separated on the sky by 10′′. The data were reduced using a custom set of IDL routines. Individual exposures were bias-subtracted and flat-fielded. The sky was then subtracted from each exposure using an optimal sky subtraction technique based on Kelson (2003), whereby the sky counts are modeled as a function of wavelength and slit – 8 – position without rectifying the original data. Object traces that were visible in a single 3000 second exposure were masked when fitting the sky. The reduced two-dimensional spectra from all exposures and both FORS2 CCDs was then combined into single array with spectral dispersion closely matching the original exposures. Nearest-neighbor sampling was used when combining the frames to avoid correlating adjacent pixels. In order to avoid false detections, hot and cold pixels, bad rows, charge traps, and other defects were identified in the flat fields and reduced frames and aggressively excluded when producing the combined spectrum. Pixels with significant dark current, identified by combining large numbers of reduced exposures, were also rejected. In total, roughly 2.5% of the illuminated area of each CCD was masked when producing the combined spectrum. Combinations of various sub-samples of the data were also made to check that no spurious features remained. The resulting 2-dimensional spectrum is shown in fig 1. A flux calibration using several spectrophotometric standard stars reproduced the published flux of the QSO to within about 20% . In the center of the final, sky-subtracted spectrum, a flux density of 1.6 × 10−20 erg cm−2 s−1 A−1 produced 1 ADU per 0.252′′ × 2′′ × 0.64 A wide pixel. The observed standard deviation near the center of the spectrum is 2.0 ADU. We can reconstruct a surface brightness profile of a line emitter by integrating over the line in the spectral direction (over one FWHM, ∼ 286 kms−1). This gives a 1 − σ surface brightness detection limit of 8.1 × 10−20 erg cm−2 s−1 ⊓⊔′′−1, if measured in a one ⊓⊔′′ aperture. The "seeing" profile as measured off the QSO near the center of the combined final spectrum is 1.07′′ FWHM. The seeing conditions generally where not as bad as anticipated, with 89% of the seeing better than 1.5′′. The large number of continuum sources on the slit, the presence of a few brighter stars and galaxies with PSF wings visible at large distances, together with charge transfer and other cosmetic problems, exascerbated by a finite number of dithering positions, made a selection of objects with an automatic method impractical. Thus, emission line objects were selected by eye. The selected depth clearly varies with the character of the object (extended, or point sources), but we estimate that for an extended object (angular extent > 1⊓⊔′′) we get a visual detection at a surface brightness 2 × 10−19 erg cm−2 s−1 ⊓⊔′′−1, reflected in the dearth of objects with a central surface brightness below that value in fig. 1. Quantitatively, this corresponds to about three standard deviations in the more precise surface brightness detection threshold based on pixel noise given above. We have tested the reality of the emitters found by splitting the sample into halves. We were able to cross-identify all but two objects (numbers 18 and 25; see below) between the halves. One more (relatively bright) object (ID 2) was entirely absent in one half but had already been excluded from the original sample because of its suspicious shape. Perusing – 9 – the combined spectrum from the Gemini telescopes we are able to cross-identify 9 of the 22 ESO objects of our emitter sample with common coverage by both the ESO and GMOS slits. However, most objects are barely detected in the shallower Gemini data. The list in table 2 contains 4 objects with a significance (based on the standard deviation of the flux) of less than 4σ. Of these, object 18 may or may not be real. 19 and 26 certainly are present in both halves of the dataset. Object 25 would not have been considered a detection everywhere in the two-dimensional spectrum but struck the eye because it seemed to constitute a group with the nearby objects 24, 26, and possibly 19 and 21. Thus we estimate that about two objects may be spurious detections. On the other hand, there are also a couple of candidates which could have been included in the sample but were not. 3. Observational Results Spurious detection is a less serious concern than misidentification and the resulting contamination of the emitter sample by foreground low redshift galaxies, as lower redshift [OII] 3727A (a doublet marginally resolved at our resolution), [OIII] 5007 A, or the HI Balmer series in emission could be mistaken for HI Lyα. We checked for the presence of these features and for others that could give away a low z object, like the absence of a Lyman α forest decrement (when a continuum was present), and the existence of extended emission in the broad band image. Among the emission line objects, five line emitters could be identified with foreground galaxies. They are listed in table 1 and shown in fig. 2. The table gives the ID numbers, redshifts, the sky-subtracted flux F in units of 10−18erg cm−2 s−1 measured in a 2′′ × 2′′ × 755 km s−1 aperture, the maximum surface brightness along the slit in units of 10−18erg cm−2 s−1 arcsec−2, and the source of the identification as a foreground galaxy. As we were interested only in line emitters we ignored the continuum-only objects for the time being. Three of the emission line objects accidentally on the slit were identified with foreground galaxies based on the presence of [OII] (two spatially almost coincident objects at z=0.39278 and z=0.4336, with IDs 7a and 7b; a galaxy at z=0.4019, ID 31). A fourth object appears to show an [OII] doublet on the very edge of the detector, and is clearly identified by several HI Balmer emission lines as a z=0.198 galaxy (ID 22). A fifth object shows Hβ and the Mg b triplet at z=0.0458 (ID 8). The contamination of the remaining sample of emitters by [OII] and [OIII] is still a concern because we do not have information about the continuum for most objects. We will treat the sample of emitters formally as Lyα in most of the paper but will return to a discussion of contamination below. The remaining 27 line emitters are listed in table 2. Fig. 3 shows the two dimensional spectra of all remaining candidate Lyα emission – 10 – line regions. Numbers correspond to the 'ID' entry (column (2)) in table 2. The 'stamps' are 60 × 60 pixels wide, with an individual pixel size of 0.252′′ × 0.64A. Thus each stamp corresponds to 15.12 ′′(116 physical kpc at the central redshift, z=3.2) in the spatial direction (vertical) and 38.4A (2266 kms−1 at the CCD center) in the dispersion direction (horizontal; wavelength increases toward the right). The spectra have been heavily smoothed in both the spectral and spatial direction (with a 7x7 boxcar filter) for display purposes and to emphasize coherent regions of extended emission. All spectra are displayed to within the same color stretch to demonstrate the variety in their appearance. Pixels within the light grey (turquoise in the color version) areas correspond to a flux density > 1.5×10−20erg cm−2 s−1 A−1. A closeup of the spectra, showing 7.56′′ by 25.6 A in a less highly smoothed (3 × 3 pixels) version is seen in fig. 4. Here the colorstretch was done individually for each image to emphasize the individual dynamic range. The QSO Lyα emission line region (ID 14) is also shown for reference but was not included in the actual analysis. Next to the 2-d spectra are the object IDs, a background-subtracted one dimensional spectrum produced by collapsing the box along the slit direction (vertical) in units of erg s−1cm−2A−1, followed by a spatial surface brightness profile along the slit direction that was obtained by collapsing the box in the dispersion direction and averaging the bottom and top sections to improve the signal to noise ratio. The velocity and spatial zeropoints are the left edge of the box and the center of the box, respectively, where the box is centered on the peak of a Gaussian fit to the spectral and spatial light profile. The 1-d spectra often look poor which is due to the faint signal and the box size, folding in a lot of noise, and sometimes wings of other objects. Optimal extraction has also been performed but with surprisingly little improvement. To help judge whether a surface brightness profile is extended, the spatial profiles to the right also show a model fit, where the surface brightness S(y) is represented by a Gaussian core plus power law wings (dotted profile). First the Gaussian with the fixed width of the point spread function (dashed profile) is fit to the innermost two pixels, to determine the amplitude parameter A for the core of the emission. Then, beyond a distance yt along the slit from the center of the emission, a power law replaces the Gaussian. The transition distance along the slit from the center of the emission, yt, and the power law index are treated as free parameters: S(y) = A exp[−y2/(2σ2)], yt(cid:17)α t /(2σ2)](cid:16) y = A exp[−y2 y < yt, , y > yt. (1) In one instance the spatial fit failed, because there were other objects nearby. – 11 – Most of the objects seem to have a relatively well defined emission peak, often sur- rounded by diffuse spatial emission, sometimes with rather broad emission lines (fig. 3). We have tried to losely classify these objects visually, according to whether they are consistent with being point sources (i.e., have a Gaussian seeing profile with a FWHM ∼ 1.07′′, as de- rived from the QSO), labelled as 'PS' in table 2; whether they are dominated by amorphous extended emission, labelled as 'EXT', or a mix between the two, centrally dominated ('CD') emission with yet an extended halo around them. The distinction is often subjective, and used only to introduce some nomenclature to aid the discussion. In this scheme, objects with IDs 1,4,9,10,15,16,17,18,20, 28, 36, 37 and 39 are clearly extended. IDs 6,9,16, 19, 23, 29, and 30 could be classified as centrally dominated. IDs 3,12,21,24,25,26,27, and 33 are in the somewhat better defined 'point source' class. Several among the brighter, mostly PS and CD sources (IDs 3, 23, 27, 28, 29, 39) exhibit the classic asymmetric line profiles known from previous studies of starbursting galaxies (e.g., Franx et al 1997; Mas-Hesse et al 2003), with a blue cutoff and a more extended red wing. A subset of those (3, 23, 28 and 29) appear to have a weaker blue emission line in addition to the red dominant one, perhaps a sign of the double humped emission profile expected from a static, externally illuminated slab (e.g. Neufeld 1990, Zheng & Miralda-Escude 2002). Intriguingly, at least one object (ID 15) shows the opposite situation, a stronger blue line opposed by a weaker red one, and there are two other bizarrely shaped objects (36, 37) where the emission seems to occur blueward of the absorption the objects cause in a nearby galaxy continuum. In those two cases the emission seems to drift redward toward larger distances from the absorber. The object 38 shows a large emission region on top of a diffuse continuum, but because of its low S/N ratio remains unclassifiable. The properties of the detected sources are listed in table 2. The columns give: (1) the number of the entry in the table; (2) the identification number of each object, as used throughout the paper (including the figures); (3) the redshift, assuming the emitter is Lyman α 1215.67 A; (4) the sky-subtracted flux F in units of 10−18erg cm−2 s−1 measured in a 2′′ × 2′′ × 755 km s−1 aperture; (5) the ratio of that flux to the one measured in a larger (2′′ × 7.6′′ × 1510 km s−1) aperture; (6) the maximum surface brightness along the slit in units of 10−18erg cm−2 s−1 arcsec−2; (7) the FWHM velocity width of a single component Gaussian fit to the optimally extracted emission line as a crude measure of overall velocity width, even where the line shape was distinctly non-Gaussian; (8) the Gaussian amplitude of the central emission region for the surface brightness model profile described in the text; (9) the turnover distance between Gaussian center and power law wings for that model; (10) the power law index for that model; (11) objects are losely classified as PS (point source), – 12 – EXT (extended) or CD (centrally dominated), and peculiarities are noted. The table shows a few instances where the ratio between the fluxes measured in the smaller and larger apertures was formally larger than unity. This happens when horizontal streaks on the CCD and sky residuals enter the larger window but not the small one, or when the background subtraction windows where in different positions for the two apertures. The errors quoted are standard deviations propagated in the usual way from the original pixel photon fluctuations. The 1σ noise of the sky-subtracted flux (entry (4)) is quoted for the 2′′ × 2′′ × 755 km s−1 aperture, the noise for the maximum surface brightness (entry (6)) is per pixel. The signal-to-noise ratio attained is totally dominated by the sky background with minor contributions from the detector noise and suppression of cosmic rays. A histogram of the wavelength distribution, together with a plot of the relative sensi- tivity of the observation, is given in fig. 5. 3.1. Spatial Profiles The size of the emitters can be characterized by a contour at which the surface brightness has dropped to a particular level. We define somewhat arbitrarily a projected "size" along the slit as the distance y from the center of the object along the slit at which the surface brightness of the fit model (1) has risen to S(y) = 1×10−19 erg cm−2s−1⊓⊔′′−1 (approximately the standard deviation in surface brightness in a 1⊓⊔′′ aperture), going inward. This is close to the surface brightness level that corresponds to a 3σ detection threshold in an opening given by the product of slitwidth( 2′′) and FWHM (1.07′′), i.e., it would be the approximate detection threshold for an unresolved source. With "size" we mean, in the following, half the total extent along the slit, analogous to the "radius" of an object. Note that this is generally an underestimate of the true radius. The two are only identical if the slit were centered on the emitter. The distribution of these "sizes" for our 27 objects is given in fig. 6. Most of the sizes occur near the spatial resolution limit (ca. 0.54′′, Half WHM), but there is a considerable tail to much larger radii. Four large objects that clearly extend beyond our fitting range are collected in the bin at 3.6′′. This bin comprises the objects 4, 15, 23, and 29. The median "radius" is 0.99′′, or 7.7 physical kpc for a Lyman α emitter at redshift 3.2. – 13 – 3.2. Number Densities and Fluxes Some of our 27 emitters are unambiguously identifiable (as HI Lyα 1216 A) just based on the line profile shapes. The same is true for the identification of the additional four, bright double component objects as [OII] 3726,3729 A emitters. Nevertheless, the low signal-to- noise ratio and the mostly invisible continuum do not permit us to apriori exclude the identification of the majority of the sources with either of those two classes of objects, 2.667 < z < 3.751 Lyα, or 0.196 < z < 0.550 [OII]. The 16 reddest of the 27 sources are at least in the right redshift range (0 < z < 0.16) to be also eligible for [OIII] 5007 A, as the weaker [OIII] transition and Hβ usually would be too faint to be seen. Although these possibilities are not equally likely, as we shall argue below, it is instructive to look at the implied number densities, luminosity functions, and fluxes for the three extreme interpretations. For the purposes of this paper we adopt a flat cosmology with Ωm = 0.3, ΩΛ = 0.7, and H0 = 70 km s−1 Mpc−1. For the following we have not applied upward corrections for slit-losses to the luminosi- ties, which occur when an object is larger than the slit. These corrections can be important, but are generally uncertain for a population of objects with an apparently large range of sizes and profile shapes. For a detailed discussion of slit losses we refer the reader to the Appendix. We emphasize here that the luminosities for objects with characteristics encountered in our sample may be underestimated by up to factors 2-5. The redshift range of the spectrum, assuming HI Lyα, is ∆z = [2.667, 3.751] = 1.085. For [OII] and [OIII] these values are ∆z = [0.196, 0.550] = 0.354, and ∆z = [0., 0.154] = 0.154, respectively. The solid angle subtended by the slit is 0.252 ⊓⊔′. The number of objects per unit redshift per square arcminute for Lyα is given by ∂2N ∂z∂Ω = 3.66 × NLyα = 98.7, where NLyα = 27 is the total number of objects. For [OII] this value is ∂2N ∂z∂Ω = 11.2 × N[OII] = 302.7, for 27+4=31 putative [OII] emitters. (2) (3) – 14 – For [OIII], ∂2N ∂z∂Ω = 25.77 × N[OIII] = 412.3, (4) with N[OIII] = 16 objects in the right wavelength range. For the case of Lyα the (cumulative) distribution ∂2N>FLyα/∂z∂Ω of those objects ex- ceeding a line flux FLyα [erg s−1 cm−2] is given by fig. 7. The comoving survey volume is given by dVc = c H0 (1 + z)2D2 A(z)dΩ dz E(z) , with the angular diameter distance DA(z) DA(z) = and c H0(1 + z) Z z 0 dz′ E(z′) , E(z) = pΩm(1 + z)3 + ΩΛ. (5) (6) (7) Then the comoving volumes represented by the two-dimensional spectrum are Vc = 885 Mpc3 h−3 70 for Lyα, Vc = 57.5 Mpc3 h−3 70 for [OII], and Vc = 1.82 Mpc3 h−3 70 for [OIII]. The total comoving number density of objects detected, dN /dVc, is then 0.030, 0.53, and 14.9 Mpc−3 h3 70, for HI, [OII], and [OIII]. The line luminosity is given by L = 4πD2 L(z)F, (8) with the line flux F and the luminosity distance DL = (1 + z)2DA, and DA as given above. The histogram of luminosities for HI Lyα is given in fig. 8. Figs. 9, 10, and 11 show the cumulative comoving density of objects versus luminosity, again under the assumption that the objects are entirely either HI Lyα, [OII], or [OIII] emitters. All luminosity functions are calculated from the fluxes measured in the larger, 2′′ × 7.6′′ × 1510 km s−1 aperture, which is related to the fluxes from the smaller 2′′ × 2′′ one (table 2, column (4)) by the factors given in column (5). – 15 – 3.3. Rate of Incidence per Unit Redshift Each population of emitters produces a total 'footprint' in the plane of the sky, that can be used to calculate the rate of incidence along a line of sight, e.g., in a QSO spectrum, once the number density and cross section on the sky are known. The contribution of the emitter population to the rate of incidence per unit redshift dN/dz, is given by dN dz = Xi σi Vi dl dz , (9) where the sum is over emitters with index i. Vi is the comoving volume in which object i could have been detected, σi the comoving spatial cross-section of emitter i, and the comoving distance per unit redshift at redshift zi is dl(zi) dz = c H0E(zi) . (10) We use the distribution of sizes (fig. 6) to compute the comoving spatial cross sections σi and plot the resulting cumulative dN/dz as a function of object size in fig. 12, where each graph is derived as if all objects where either entirely HI Lyα, or [OII]3727, or [OIII]5007 emitters. The short vertical lines denote the resolution limit, telling us that objects with nominal sizes smaller than that may not be contributing as much to the cross-section on the sky and thus to the dN/dz. This correction is relatively small for all cases. 4. The Identity of the Emitters Most of our objects are technically single line objects, too close to the detection threshold to study the precise line shape or detect other weaker lines that should also be present (as in the case of [OIII]). Moreover, for most of them neither the spectrum nor a deep Keck LRIS V band image (see below) show significant continua. Judging from the emission line profiles alone, we estimate conservatively that at least about six objects are likely to be high redshift Lyman α because of their pronounced asym- metric emission profile. A further three objects seem to coincide closely with QSO Lyα forest absorption systems (see below), which makes them relatively secure HI identifications. Thus, from the spectroscopic evidence discussed so far it is not possible to exclude that the majority of our objects are low redshift contaminants from [OII] or [OIII]/Balmer series – 16 – emitters. We shall now present a number of arguments that will help to judge the plausibility of these alternatives. 4.0.1. Are the emitters dominated by [OII] at low redshift ? In addition to the 4 objects already eliminated based on their bright [OII] doublets and continua (see section 3) there is a similarly small number among the 27 remaining objects the line profiles of which seem at least to be consistent with [OII]. We can only barely resolve the [OII] 3736,3738 doublet in our spectra, but we estimate that we see up to four objects with potential multiple emission peaks (probably including noise spikes) at the right wavelength separation, that are at least consistent with [OII] (though none of them has to be). Given our rather poor spatial resolution of 4-8 kpc at these redshifts we should probably expect any additional [OII] emitters to reside amongst our spatially unresolved sources. If our emitters were [OII], the luminosities would range from 3.7×1038erg s−1 to 2×1040 erg s−1 and the redshifts range from z = 0.196 to z = 0.550. Using a standard calibration for the relation between star formation rate and [OII] luminosity, SFR(M⊙yr−1) = 1.4 × 10−41L[OII](erg s−1) (11) (Kennicutt 1998), these luminosities correspond to star formation rates between 5 × 10−3 and 0.3 M⊙ yr−1. The total inferred [OII] luminosity density of 1.0 × 1041 erg s−1 Mpc−3 would correspond to a star formation rate density of 1.42M⊙yr−1Mpc−3. The space density correponds to 0.53(N/31) Mpc−3. We can estimate the number of expected [OII] detections from the field galaxy luminosity function of Trentham, Sampson & Banerji (2005). We are able to detect emitters with line fluxes > 10−18erg cm−2s−1. Most [OII] emitters (Hogg et al 1998) do not exceed a rest equivalent width of 50 A. If we use this value to convert our line flux detection threshold into continuum magnitudes, and adopt the redshift 0.364 that divides the volume where we can detect [OII] into a lower and higher redshift half, we find M = −12.8 as the faintest continuum magnitude where we would be able to detect the corresponding [OII] emission. The total space density of local field dwarf galaxies down to an absolute magnitude of MR = −13 is about 8 × 10−2Mpc−3 (Trentham, Sampson & Banerji 2005). Based on that estimate, about five of our emitters are indeed likely to be due to [OII] emission in low redshift dwarf galaxies. This number fits in well with the four [OII] emission line galaxies which we have already identified, outside of the 27 faint emitters. Just from Poissonian arguments (again assuming that we are not looking at a cluster), the probability to have 2 more additional [OII] emitters in our volume is about 22 %, the probability to have more than 4 more is only about 7 %. – 17 – The rate of incidence of low redshift damped Lyα systems is another (albeit somewhat uncertain) indication that our emitters are not predominantly [OII]. If we assume that all star-forming galaxies are embedded in DLA (are a subset of DLAs), and that the radius of optically thick gas is very likely to be larger than the one of [OII] emission, the product of number density and cross-section of [OII] emitters cannot be larger than that of DLAs. Our inferred total dN/dz (fig. 12) for [OII] at 0.93 is more than 14 times larger than that of its contemporary damped Lyα systems, estimated from the Sloan Digital Sky Survey, dN/dz(z = 0.37) = 0.066 (Rao, Turnshek & Nestor 2006), and so even if DLAS were not larger than [OII] emission regions, and not more numerous, only about 7% of our emitters (i.e. two objects in total, or none in the remaining sample of 27) should be [OII] to not violate the dN/dz constraint from DLAS. 4.0.2. Are the emitters dominated by [OIII] at even lower redshift ? As for [OIII] 5007A, there is a total of 16 emitters in the redder 769 A long part of the spectrum where [OIII] could be detected. The remaining 12 emitters occupy the bluer 550 A, where the wavelength is below the rest frame wavelength of [OIII] and obviously cannot be [OIII]. The ratio between the numbers of objects per wavelength in the [OIII] region versus those in the non-[OIII] region is then 0.95 ± 0.23, i.e., there is no significant enhancement of the line density, making a dominant contribution from [OIII] emitters unlikely (a similar line of reasoning can be employed against the emitters being [OIII] 4959 A etc). The corresponding luminosities would range from 2.0× 1036 erg s−1 to 3.2× 1038 erg s−1 at redshifts between z = 0 and z = 0.154. The space density corresponds to 9(N/16) Mpc−3, about 40 times higher than the local space density of dwarf galaxies down to an absolute magnitude of MR = −9 (0.23 Mpc−3; Trentham et al 2005). As we have found already one galaxy in the right redshift range outside of our emitter sample (though one that had only Hβ and no actual [OIII] emission, see section 2), the Poissonian probability to have one or more additional ones hidden in our sample is less than 7%, the probability to have two or more less than 1%. There is no obvious foreground cluster in our field and it seems thus very unlikely that even one of these emitters is due to [OIII] emission from an HII region in low redshift dwarf galaxies. Note, however, that curiously the lower end of the inferred luminosities at the smallest distances corresponds to that of bright planetary nebulae (PNe; 4 objects). Gerhard et al. (2005, 2007) have searched for PNe in the core of the Coma cluster with a multiple slitlet technique to similar limiting fluxes. They found 35 PNe candidates in a similarly-sized volume but centered on the core of the Coma cluster, where the overdensity of PNe should be very large. It appears thus unlikely that in a random field we should have – 18 – found [OIII] emission from bright PNe. The rate of incidence of DLAs (at mean redshift 0.08) constrains the fraction of [OIII] emitters, too, limiting it to be less than 15% of our emitters. We conclude that we are likely to have found (and eliminated already) most of the [OII] contaminants in our emitter sample, as predicted by the space density of the local galaxy population, and the rate of incidence of low redshift damped Lyα systems. It is even less likely that there are [OIII] 5007 A contaminants in our sample. Therefore, from here on we shall treat the remaining emitters as HI Lyα and discuss the implications, keeping in mind that some of the objects may still be misidentified. 5. Lyman α Emitters at Redshifts 2.666 < z < 3.751 If our objects are Lyα emitters the observed fluxes correspond to luminosities between 7.9 × 1040 erg s−1 to 1.6 × 1042 erg s−1. If caused by star formation, the range of luminosities corresponds to star formation rates of 7 × 10−2 to 1.5 M⊙ yr−1 , where we have used the standard relation SFR(M⊙yr−1) = 9.1 × 10−43LLyα(erg s−1) (12) for the Lyα luminosity as a function of star formation rate (based on Kennicutt (1998) and case B assumptions for the conversion of Hα and Lyα ; Brocklehurst 1971). Note again that the actual values could be larger by a factor of a few due to slit losses. The total Lyα luminosity density of 1.4× 1040 erg s−1Mpc−3 corresponds to a star formation rate density of 1.2× 10−2M⊙yr−1Mpc−3, about 36% of the value (uncorrected for dust) inferred for B-band dropouts in the Hubble Ultra Deep Field by Bouwens et al 2007. The inferred space density is 3.0(N/27)10−2 Mpc−3, a factor three smaller than the total space density of local dwarf galaxies, but an order of magnitude larger than the space density of previously known Lyα emitters at this redshift (but with our study going down to much lower flux limits). If our detected emission line objects are primarily due to Lyα this correponds to a significant steepening of the luminosity function of Lyα emitters at luminosities below ∼ 1042 erg s−1. Is such a numerous population of Lyα emitters plausible? The inferred space density is similar to the space density of B droup-outs in the HUDF at slightly larger redshifts (Bouwens et al. 2007) (see also below, and fig. 15), and in fact, less by factors 10 − 30 than the number density inferred by Stark et al (2007) for z ∼ 8 − 10 objects. Intriguingly, – 19 – the abovementioned survey for planetary nebulae by Gerhard et al (2005) also found 20 "background objects" that could not be identified otherwise, in a similar volume. Other dedicated surveys for Lyα emitters at z ∼ 2.5 − 3.5 (e.g., Hu, Cowie & McMahon 1998; Kudritzki et al 2000; Steidel et al 2000; Stiavelli et al 2001; Fujita et al 2003; van Breukelen et al 2005, Gronwall et al 2007, Ouchi et al 2007) appear consistent with our survey (there is little overlap in the range of fluxes reached). Our objects have about 20 times the volume density of, for example, the objects found by the shallower Gronwall et al survey (their detection threshold is 1.5 × 10−17 erg cm−2 s−1). A comparison of our cumulative luminosity function (solid line, with 1σ errors indicated by the dotted lines) with the best fit, z=2.9 luminosity function from the IFU survey by van Breukelen (dash-triple-dotted line), the z=3.1 (narrow band filter) luminosity function of Ouchi et al 2007, and the predictions by LeDelliou et al (2006, dash-dotted line) is shown in fig 13. Note again that our luminosity function does not include corrections for slit losses. At the bright end of our sample (near ∼ 1042 erg s−1) there is good agreement with the two observed luminosity functions, and there is initial agreement in the slope as well in the small region of overlap, but our sample becomes steeper going toward fainter magnitudes. The solid line appears to flatten again toward luminosities below 2 × 1041, (or a flux 3 × 10−18 in the usual units). Objects at half that flux are still clearly detectable for sources with characteristics similar to ours. This suggests that the flattening may be real, and the numbers may start to decline. One possibility is that we may already be seeing the bulk of the currently star forming galaxies. From a theoretical point of view the CDM picture of structure formation predicts a rather steeply rising mass function at low masses, but note that even for a linear light- to-total halo-mass relation the luminosity function required to explain our inferred space density requires an even steeper faint end slope. Near 1042erg s−1, our observed density of objects is in agreement with the CDM based model population of Lyα emitters from LeDelliou et al. (2005, 2006; dash-dotted line in fig 13), but then it steepens over the next decade in luminosity considerably, relative to the models based on a constant Lyα escape fraction, perhaps suggesting that this fraction may not be constant after all. The discrepancy approaches about a factor 5 at 2 × 1041erg s−1 and then decreases again toward fainter magnitudes. Such a steepening of the luminosity function could perhaps be explained if dust extinction becomes increasingly less important for fainter emitters. Unfortunately, the stellar and total masses of the emitters are very uncertain. If the Lyα emission is due to star formation at the rates estimated above, the accumulated stellar mass within 109yr is in the range 7 × 107 M⊙ to 1.5× 109M⊙. Another estimate of the mass can be obtained by comparing our inferred space density with that of dark matter halos – 20 – predicted by CDM models. The space density inferred by our sample of objects corresponds to the cumulative space density of dark matter halos with total mass > 3 × 1010 M⊙ and circular velocities vc > 50kms−1 (e.g. Mo & White 2002, Wang et al. 2007). In the next chapter we shall examine the competing Lyα production mechanisms, be- fore returning to a discussion of the nature of the emitters, in the larger scheme of galaxy populations. 6. Astrophysical Origins of the Lyα Emission At the faint detection threshold attained here a number of different physical processes can produce Lyα emission at comparable fluxes, and it is not certain that we are neces- sarily seeing the result of star formation. The faintest of these competing mechanisms is Lyα fluorescence, induced by the general UV background. However, our fluxes (see table 2) typically exceed the predicted surface brightness limit for individual objects (Gould & Weinberg 1996) by an order of magnitude. A second source of Lyα photons arises from the presence of a QSO in our field. The QSO locally enhances the UV flux and can in principle boost the surface brightness of HI to much higher levels where it can be readily detected. A third effect expected to rear its head at our sensitivity threshold is cooling radiation; gas falls into a galactic potential well and sheds part of its potential energy in the form of Lyα line radiation. These processes are observationally distinct from star formation in that the latter is the only one that actually produces a significant (stellar) continuum as well, which can serve as a discriminant among the various sources of Lyα. This question will be addressed briefly in the next section, followed by an investigation of the role of the QSO's local radiation field, and of cooling radiation. 6.1. Stellar Continuum Emission from the Line Emitter Sample To check for continuum emission we could avail ourselves of V band images taken with the LRIS instrument on the Keck I telescope, with a total exposure time of 5610 s. The combined image was flux-calibrated with the photometric data from Hall et al (1996). The slit coordinate system was mapped onto the two-dimensional image and the V band fluxes were measured in appropriately positioned apertures of size 2′′ × 2′′. These apertures are expected to have a typical spatial uncertainty on the order of half a slit width perpendicular to the slit, as the spectrum allows us only to derive the coordinate along the slit. The 1σ detection threshold in this aperture is 3 × 1027h−2 70 erg s−1Hz−1. Detected V – 21 – band luminosities (crosses with error bars) together with 3-σ upper limits for undetected objects (arrows) are shown as a function of the Lyα luminosity in fig. 14. Objects 1,2,3 and 39 fell off the edges of the V band image and where not constrained. However, object 39 has a detectable continuum in the spectrum itself, and shows a clear Lyman α forest Its data-point in the plot gives the 1500A continuum flux measured directly decrement. from the spectrum. As for the other objects, visual inspection shows that very few objects selected by the presence of Lyα emission in the spectrum show up in the V band image. At the 3σ flux level, only two objects have automatically detectable V-band counterparts, namely 9, and 33, both of which could be low redshift continuum sources or high redshift line emitters experiencing chance coincidences with lower redshift continuum sources. Interestingly, these are the same two objects picked out by eye as having clear continuum counterparts. In the spectrum itself, several objects coincide with apparent continuum traces (all very faint), many of which, with the exception of the above mentioned object 39, are consistent with bad rows or charge transfer problems, or accidental spatial coincidence with unrelated continuum sources. It is instructive (and sobering) to consider where in the continuum - line luminosity diagram (fig.14) star forming galaxies should reside, were we able to detect them both in continuum and line emission. Adopting again eqn. 12 for the Lyα luminosity as a function of star formation rate, and LU V (erg s−1Hz−1) = 8 × 1027 SFR(M⊙yr−1) (13) for the UV continuum luminosity (for a Salpeter IMF, and solar metallicity; Madau, Pozzetti & Dickinson (1998)), we equate the star formation rates in these relations to obtain the dashed line in the bottom RHS corner of fig. 14. It delineates the positions of galaxies where both UV continuum flux and Lyα line flux are entirely due to star formation, and is given by log(LU V ) = −14.14 + log(LLyα). (14) The Lyα restframe equivalent width formally implied in equation (14) is 68 A, a value high for color-selected galaxies (Shapley et al 2003) but not exceptional even for much brighter Lyα emitters (e.g., Gronwall et al 2007). There is also an upper diagonal dashed line, showing the locus for a rest equivalent width of 20 A. The dotted vertical line to the left gives the Lyα flux for a star formation rate of 1/10 M⊙yr−1, the RHS one for 4/10 M⊙yr−1. Unfortunately, most of our objects are predicted to be too faint in the continuum to be able to test whether star formation is the origin of the Lyα, so the continuum flux – 22 – measurement or, equivalently, the equivalent widths are not helpful here. The non-detection of the continuum is of course fully consistent with star-formation induced Lyα emission. The position of object 39 so far (to the left) of the SF locus suggests that the Lyα emission is heavily suppressed, e.g., by dust, as seems to be the case for massively starforming galaxies (e.g., Shapley et al 2003). The situation is summarized in fig. 15, where we compare the cumulative UV continuum luminosity functions of Steidel et al (1999; asterisks) and the Hubble Ultra Deep Field z ∼ 4 B-band dropouts (Bouwens et al 2007; dashed line) with the line emitters in our survey (solid line). For all but object 39 (where we have an actual measurement) we have assigned "continuum magnitudes" based on eqn. (14). Note that this is just a scheme to show the predicted continuum if both UV continuum and Lyα line radiation would be entirely due to star formation, ignoring any extinction effects. Given our small survey volume, only about one object in our entire survey should be bright enough to have shown up in a ground-based, broad-band-color survey, i.e., as a "Lyman Break" galaxy (Steidel et al 1999), and this is what we find (namely number 39). Object # 39 brings up the number of galaxies brighter than -20.3 AB magnitudes to unity, virtually identical to the prediction from the integrated continuum luminosity functions for a volume of our size. Our volume is too small to have a much brighter galaxy in it. The total number density of our emitters is comparable to the number density of the Bouwens et al study at magnitudes brighter than MAB = −16.5. The "luminosity function" for the line emitters appears to be steepening between -20 and -18, a behavior already seen above in our comparison with the Lyα emitters. It could be indicative of dust extinction decreasing towards fainter magnitudes. A correction for dust would reduce the slope of the line emitter luminosity function, bringing it into better agreement with the Bouwens et al curve. We caution, however, that our objects cannot be not strictly identical to the class of B-band dropouts as half of them are at lower redshift. 6.2. Lyα fluorescence induced by the QSO ? We turn next to the possibility that the Lyα radiation arises from patches of optically thick hydrogen gas, induced to Lyα fluorescence by the ionizing radiation from the QSO in our field. The basic idea is that the partial conversion of the QSO UV radiation field into Lyα photons at the surface of optically thick hydrogen bodies raises the emission from clouds or galaxies in the QSO vicinity above the detection level. This effect has been studied by a number of authors (Fynbo et al. 1999, Francis & Bland-Hawthorn 2004, Francis & – 23 – McDonnell 2006, Adelberger et al 2006, Cantalupo et al 2007). The spatial extent of the zone of influence is obviously inversely proportional to the square root of the intensity of the fluorescent emission. Following Cantalupo et al 2005 we can express the enhanced Lyα flux in terms of a boost factor, i.e., the ratio of the surface brightness enhanced by the QSO (S) to the one caused by the general UV background (Sbg), S Sbg = (cid:0)0.74 + 0.5b0.89(cid:1) , where b = 15.2 × LLL 1030ergs−1Hz−1 0.7 α (cid:18) r phys.Mpc(cid:19)−2 . (15) (16) Both the luminosity of the QSO at the Lyman limit, LLL, as well as its precise systemic redshift are critically important ingredients in this calculation. Estimating the latter from the position of the OI λ1302 A emission line we determine the QSO systemic redshift as zem = 3.32209. The luminosity per unit wavelength is given by L(λ/(1 + zem)) = 4π(1 + zem)D2 L(z)f (λ), (17) With f (1040A) = 2.2 × 10−17erg cm−2 s−1A−1 measured directly from our fluxed QSO spectrum we arrive at a luminosity per unit wavelength L(1050A) = 9.405× 1042erg s−1A−1. To measure the number of HI ionizing photons we still need to determine the luminosity at the Lyman limit and the power law dependence for wavelengths below the ionization threshold. According to the study by Scott et al (2004), the power law index for a QSO with log λL(1050A) ≈ 9.8 × 1045 erg s−1 where L(ν) = L(ν0)(cid:18) ν ν0(cid:19)α (18) is (statistically) consistent with α ∼ −1.5 ( in good agreement with α ∼ −1.57 for the radio-quiet sample from Telfer et al 2002). Extrapolating the luminosity from 1050 A to the Lyman limit with L(λ) = L(ν0)(cid:18)λ0 λ (cid:19)α+2 c λ2 0 (19) we get – 24 – L(ν) = 2.798 × 1030(cid:18) ν νLL(cid:19)−1.5 erg s−1Hz−1. Inserting these results in the above relation for the boost factor gives r = 3.018 (S/Sbg − 0.75)0.5618 phys.Mpc. (20) (21) For the relatively faint QSO DMS2139.0-0405 to boost the surface flux from the back- ground value Sbg = 3.67×10−20 erg cm−2s−1⊓⊔′′−1 to a typical surface brightness of S ≥ 10−18 erg cm−2s−1⊓⊔′′−1 as observed would require the object to be within only 0.479 proper Mpc or 2.07 comoving Mpc. This distance corresponds radially to 254.8 pixels spatial pixels along the slit (about 1/4 of the length of the field), but only 4.3 (!) pixels in the dispersion di- rection. Fig. 16 shows the highly excentric elliptical contour within which to expect the enhancement to 10−18 erg cm−2s−1⊓⊔′′−1. Only one object, ID 16, falls within the ellipse, and with its maximum surface brightness of ∼ 10−18 and absence of a continuum is consistent with fluorescing in the ionizing field of the QSO. The overwhelming majority of our sources, however, appear to be oblivious to the QSO's proximity. 6.3. Signs of Radiative Transfer, and Cooling Radiation The spectral line shapes and sizes of our emitters suggest that the Lyman α photons may have been processed by radiative transfer through an optically thick HI medium. The trapping by and protracted escape of line radiation from such a medium should lead to random walk in the spatial and frequency domain. The result may be observable as emission broadened in frequency space and extended in the spatial direction beyond the extent of the actual source of Lyα photons (e.g., Adams 1972; Neufeld 1990; Zheng & Miralda-Escud´e 2002; Dijkstra et al 2006a; Tasitsiomi 2006). The data appear to show some evidence for these mechanisms at work. The large velocity widths (see table 2) and radial extent (median projected radius along the slit 7.7 kpc proper, and considerably larger in individual cases) that we have observed are thus suggestive of the signatures of radiative transfer. The FWHM velocity widths measured from optimally extracted spectra of the individual emission line regions are plotted in fig. 17 versus the power law index of the surface brightness model (equation (1)). In some cases these widths are underestimates because only a single peak was fitted, as opposed to a double humped or more complex structure. In that plot, the area to the left of the spectral resolution, about 286 km s−1 FWHM, is visible as a zone of – 25 – avoidance, and between a third and half of the measured velocity widths clearly exceed the resolution. Sources with large velocity widths seem to prefer smaller power law indices, i.e., spatial surface brightness profile that drop less rapidly with radius. 6.3.1. Spatial Surface Brightness Profiles and Fluxes To learn more about the topography of the Lyα source and the origin of the radiation we can attempt to compare our average measurements of the sizes, peak surface brightness, and total fluxes to the models by Dijkstra et al (2006). With the number of free parameters and the simplifications in these models and the observational complication of the long-slit technique it is difficult to make a quantitative comparison, but we can at least check whether the observables agree at an order of magnitude level. As far as we can tell given the limited spatial resolution, our typical surface brightness profile requires that the sources are at least somewhat centrally concentrated, similar to the model 4 of Dijkstra et al (see the surface brightness profile in their fig. 5). Our median observed "radius" (= half the extent along the slit) at the 10−19erg cm−2 s−1 ⊓⊔′′−1 surface brightness contour is about 1.0′′, the median power law slope αmed = −2.0. Making appropriate corrections for the slit losses and distortions of the surface brightness profile we find good agreement with the surface brightness profile shape of Dijkstra et al if we scale down their total flux to 1.1× 10−17erg cm−2 s−1. The mass dependence of the flux for their model 4 at z=3.2 is 4.16 × 10−18(Mtot/1011M⊙)5/3erg cm−2 s−1. Our corrected median flux 1.1 × 10−17erg cm−2 s−1 would then correspond to a cooling halo with total mass ∼ 1.8 × 1011M⊙. Thus, the median surface brightness profiles and total fluxes observed appear broadly consistent with the Lyα arising predominantly as cooling radiation. The typical halo mass required to produce the luminosity function of our emitters, is, however, uncomfortably large for cooling radiation to be the dominant source of Lyα for a majority of our objects. Dijkstra et al. (2006) estimate the expected cooling radiation assuming that the gas in DM halos cools on a free-fall time scale and cooling is predominantly by Lyα emission. The Lyα emission is then a strong function of the virial velocity of the halo, LLyα ∼ 1.6× 1039(vc/35 kms−1)5 erg s−1. Note that the free fall time scale is shorter than the time that corresponds to the redshift interval 2.667 < z <= 3.751, and a newly collapsed DM halo would only emit for about 40% of the redshift range where we can observe it. Even if we assume that all DM halos present at the lower end of the redshift interval have collapsed and started cooling in our redshift interval, all halos with vc > 35 kms−1 would be necessary to account for the observed space density of emitters. The Lyα luminosity for the typical object would generally be more than a factor ten lower than we observe even if we neglect the slit losses. This does not preclude – 26 – that the flux of a few of our emitters in more massive halos is dominated by Lyα cooling radiation but it is very unlikely that this is a large number. Lyα radiation powered by star formation appears to be the energetically most favorable explanation for the majority of our emitters. 6.3.2. Evidence for Radiative Transfer Mechanisms from Spectral Line Profiles Irrespective of the origin of Lyα photons, radiative transfer of line photons from a central source within an optical thick halo should have observational signatures characteristic of the kinematics of the gas. Several of our objects (ID # 3, 12, 21, 23, 28, 29, and 39; see figs. 3 and 4) exhibit strong, spatially concentrated emission peaks (even though their emission often extends further out) with asymmetric line profiles showing a steep drop in the blue and an extended red shoulder; such profiles have been seen previously in low- and high redshift star-forming galaxies and are generally considered to be consistent with radiative transfer in the expanding supershells of galactic winds (e.g., Lequeux et al 1995; Mas-Hesse et al 2003) At various stages of their evolution the line profiles may resemble single emission line peaks, PCygni profiles, or double component profiles with a dominant red component (Tenorio-Tagle et al 1999; Ahn et al 2003, Ahn 2004). Several of these asymmetric emitters (3, 23, 28, and 29) show a weaker blue peak oppos- ing the red one, which may be evidence for a wind shell or more generally radiative transfer through an expanding optically thick medium (Zheng & Miralda-Escud´e 2002; Dijkstra et al 2006a,b; Tasitsiomi 2006). 6.3.3. Individual Candidates for Emitters dominated by Cooling Radiation Most other emission profiles look amorphous and defy classification because of the low signal-to-noise level, but there is a small group of emitters (IDs 15, 36 and 37; fig. 18) fortuitously projected near the QSO trace, that shows a number of intriguing properties different from those of the other sources. The three continuum traces visible in that figure are the QSO (with four strong Lyα forest absorption lines of rest equivalent widths 2.0 A (A), 1.3 A (B), 1.2 A (C), and 0.9 A (D); an unrelated, featureless, presumably low redshift continuum object just above the QSO trace; and further up a faint high redshift object from which two faint emission smudges (36 and 37) seem to protrude. Even though their appearance seems unusual, the identification of all three smudges with HI Lyα is relatively – 27 – secure because of their close alignment in redshift with QSO absorption systems: the dip in the emission region of object 15 and the blue starting point of the emission region of object 37 both coincide to within less than 100 kms−1 with the strong absorption line C in the QSO spectrum between them. The projected transverse (here: vertical) distances from the QSO are 150 (object 37) and 90 (object 15) physical kpc. The object 36, at a similar transverse distance from the QSO as 37, also coincides closely in redshift with another QSO absorption system (B). Object 15, below the QSO trace, appears to consist of a strong blue emission component, separated by a dip in flux from a fuzzy redder emission bit, which also may be rotated slightly. Both, the velocity shift between the blue emission line and the dip, and the FWHM of the blue peak each amount to approximately 200 kms−1. It is difficult to be sure of what we are seeing here (perhaps two merging protogalactic clumps), but the signature of strong blue peak, central dip, and weak red peak is not unlike the one expected for cooling radiation from gas falling into a galactic halo. If this is what we are are seeing, then, according to the simulations by Dijkstra et al (their fig. 8) such relatively small values for blueshift and FWHM of the blue peak may indicate a cooling halo with relatively small infall velocities and HI optical depths. Object 37 (cf. figs. 3 and 4) shows a bar-like emission region projecting out at about a 60 deg angle on one side from the blue edge of a strong absorption line in the nearby, faint background galaxy, as if a "door" had been opened anti-clockwise in the continuum of that galaxy. The transverse extent of the emission is at least about 3.3′′or 26 physical kpc. The spectral width of the tilted emission bar is about 320 kms−1 (i.e., it is possibly unresolved), and it projects out from the continuum object, starting about 490kms−1 blueward of the centroid of the absorption line (rest EW=3.9 A) in the faint continuum object, shifting to the red with increasing distance from the continuum object by between 280 and 470 km s−1 (the uncertainty arises from the difficulty of estimating the spatial extent of the emission region). Object 36 is another broad smudge of emission, losely (the S/N is poor) lining up in redshift with QSO absorption system B (a weak absorption feature appears in the continuum object, about 110 kms−1 blueward of the QSO absorber). Again, going outward from the continuum object the emission can be traced spatially to a similar extent as object 37, but in this case extends over a larger wavelength range, becoming redder by up to 1500 kms−1. The feature is clearly resolved in velocity, with a width of about 1000 kms−1. All three objects show emission blueward of the absorption centroids of either the QSO absorption lines or the absorption in the continuum emitter. In addition, the color gradient from red to blue when approaching the absorption systems could be understood in terms of – 28 – infalling halo gas, that accelerates and cools when approaching smaller radii, as described by Dijkstra et al. However, it is not clear that the absorption systems really represent the centroids of the halos and do not rather arise in the outskirts. A scenario with out- rather than inflows and a different topology cannot be excluded, at least not for objects 36 and 37. There are two more QSO absorbers in that group that span a total (from A to D) of 52.75 h−1 Mpc (or 20 physical Mpc), if the redshift difference is due to the Hubble flow. The fact, that the three most unusual emitters, together with a cluster of strong QSO absorption lines occur in a spatially relatively narrowly but apparently highly elongated region (even the line of sight distance between B and C is 7 physical Mpc (or 18.43 h−1 comoving Mpc) suggests that we may be looking along a large scale filament or sheet, with the QSO absorbers representing the outskirts of the three galaxies whose emission regions we see. 7. Correspondence between the Line Emitters and Optically Thick Lyα Forest Absorption Systems The existence of two independently identified classes of optical thick objects in the universe, Lyα emitters and Lymit Limit absorption systems, enables us to establish a corre- spondence between them and constrain their properties. If we assume that the emitting and absorbing regions are identical in size (in reality, the absorption cross section may be an upper limit to the emission cross section, for optically thick gas), we can equate the rate of incidence per unit redshift, dNLL/dz, of Lyman limit absorbers above a certain HI column density, and the product of the spatial comoving density of Lyα emitters, their emission cross-section, and the redshift path, where dNLL dz = Xi σi Vi dl dz , dl(zi) dz = c H0E(zi) . (22) (23) The comoving density of objects is again dN/dVc = Nobj/(885h−3 70 Mpc). From fig. 19, our total observed rate of incidence of Lyα emitters would be dN/dz = 0.30, if all 27 objects were Lyα emitters. They same cautions as mentioned above about slit losses apply to the estimate of the radius from the size along the slit. If the emitters had a spherical, sharp-edged outline in – 29 – the plane of the sky, then by adopting the total extent along the slit for the diameter of the object we would underestimate the latter by a factor π/4. The finite sizes of the objects also affect the detectability on the slit, as an extended object may be detected even if its center falls outside the slit. This increases the effective comoving volume and decreases the space density for the objects. We use a simple model to compute the comoving volume, where an object of a given radius can be detected if part of it fills the slit. Our correction can only be indicative of the true corrections. The little we know about the emitters at present does not warrant a more detailed approach. The resulting dN/dz, corrected for these effects, is shown in fig. 19 as the dotted line. The correction emphasizes the relative contributions to dN/dz from objects smaller than the slitwidth and reduces the relative contribution from larger objects. The total correction (larger cross-section and smaller comoving volume) reduces the overall dN/dz by about 22%. to dN/dz = 0.23. About half of the contribution to dN/dz arises from the four most extended objects. Interestingly, the observed dN/dz for our emitters and the one for DLAS with neutral hydrogen column densities NHI > 2 × 1020cm−2 ( dN/dzDLA = 0.26; Peroux et al 2005; Storrie-Lombardi & Wolfe 2000) are comparable. In other words, the combination of large sizes and the high space density of the emitters together imply that the total rate of incidence is sufficient to explain the majority of damped Lyα systems. It thus appears that we may have finally detected the star formation associ- ated with most of the rate of incidence of damped Lyα absorption systems in the redshift range 2.667 < z < 3.751. The low star formation rates of 0.07 to 1.5 M⊙ yr−1 would explain the low success rate of direct searches for the host galaxies of DLAS. If the interpretation of the emission as being due to the hosts of DLAS is correct, then we have for the first time established the typical size and space density of DLA host galaxies. Within the CDM model of structure formation we can then also infer their masses (3 × 1010 M⊙ total, 5 × 109 M⊙ in baryons) and virial velocity scale (∼ 50 km s−1). The typical values for size, mass and virial velocity agree well with the predictions of the model of Haehnelt, Steinmetz & Rauch (1998, 2000), who interpreted the observed kinematic properties of the neutral gas in DLAS as probed by low-ionization species within the context of CDM models. The inferred star formation rates and the inferred masses are consistent with the low observed metallicities of DLAS which are generally overpredicted in models assuming larger star formation rates. Note that if all the observed emission is due to Lyα, the total star formation rate density corresponds already to 36(fslit/1.0) percent of the total non-dust-corrected star formation rate inferred from drop-out studies at these redshifts (e.g. Bouwen et al. 2006; fslit is the factor by which the observed flux needs to be multiplied to correct for slit-losses). Unfortu- – 30 – nately, we have no direct observational handle to decide whether the star formation itself is extended. If all the emission is Lyα, our total star formation rate density is, however, higher by at least an order of magnitude than the upper limits obtained by Wolfe and Chen who searched for extended continuum emission from DLAS in the HUDF. At the same time, our star formation density is close to (about 60% of) the value needed to explain the heating of DLAS (Wolfe, Gawiser & Prochaska (2003)). As suggested by Wolfe et al these discrepancies can be reconciled if the star formation in question is confined to a compact region at the cen- tre of DLA hosts, rather than arising in large stellar disks. The Lyα emission in our objects often appears extended, but this does not necessarily mean that the sources of the ionizing photons responsible for producing the Lyα photons are similarly large. The extended nature of the Lyα emission would be due to resonant line scattering, with Lyα photons random walking their way out to radii that have never seen a star. Alternatively, some of the large sizes could be due to unresolved merging protogalactic clumps, a situation that is common in a CDM scenario and may explain the observed kinematics of DLAS (Haehnelt et al. 1998). Finally, the identification of the emitters with DLAS, which are known to be essentially dust-free (e.g. Murphy & Liske 2004), and the realization that Lyα from emitters brighter by an order of magnitude have Lyα emission reduced by a factor 3, presumably by dust (e.g., Gronwall et al 2007), would also explain at least in part the steep rise of the number of emitters over a decade in surface brightness as a drop in dust content. 8. The Hogan-Weymann Effect Sofar we have not addressed the effect of Lyα fluorescence, induced by the general UV background (Hogan & Weymann 1987). The original expectation was that the number of objects lit up by the UV background corresponding to optically thick LLS would yield about the same number of objects (30, with a radius of 2.5′′ at column density 1019 cm −2) as we have found, albeit at considerably lower surface brightness. A possible explanation is that many LLS may have ongoing local, low level star formation. This would be consistent with LLS having somewhat higher metallicities (Steidel 1990) than the general Lyα forest (Simcoe et al 2004). Given our surface brightness profiles it is possible that the underlying fluorescence is simply swamped by star formation Lyα. There are several conceivable approaches to searching for low light level emitters in the field. Originally we had considered the possibility of a blind search (e.g. Bunker et al 1998), which, however would have reached only a sensitivity a factor two above the recently revised (lower!) estimates for the anticipated signal, based on the opacity of the Lyα forest (Bolton et al 2005). One possibility to increase the sensitivity, suggested by the number of – 31 – objects already detected is to search in their immediate neighborhood for an extended signal of diffuse emission surrounding the brighter star forming regions that is in agreement with the expected surface brightness (Gould & Weinberg 1996) 9 × 10−20(cid:16) η 0.5(cid:17)(cid:18) J 4.3 × 10−22(cid:19) erg cm−2s−1⊓⊔′′−1, (24) at < z >= 3.2 for a QSO dominated, Haardt & Madau 1996 UV spectrum with slope 1.73; η is the fraction of the energy of the impinging UV photons converted into Lyα, and J is the UV background. The value for J in equation (24) is scaled to the recent measurement of the photoionization rate at z ∼ 3 in the Lyα forest (Bolton et al 2005). An ionizing spectrum in which galaxies and QSOs each produce half of the ionizing flux would result in a surface brightness lower by a factor 2 than the fiducial value in eqn. (24). Our observation is deep enough to probe the surface brightness for general fluorescence if we combine the signal from all sources to improve the signal-to-noise ratio. Boxes of spatial width 2′′ and spectral length 1500 km s−1 (to include the likely extent of the double humped emission profiles predicted; e.g., Cantalupo et al 2005) have been placed on either side of our emitters, at varying distances along the slit direction. The boxes were sky-subtracted once more locally, using windows below and above (along the slit) the box used for the signal, but further out from the emitter than the signal boxes. Because of the strong presence of weak continuum objects in the 2-D spectrum only a subset of the emitters (usually 12-14) were in a sufficiently clean area of the field to be useful for this analysis. The mean and median surface brightnesses and their statistical errors were extracted. The results are presented in fig. 20. The open squares show the medians, and the crosses with error bars give the total weighted mean surface brightness of the boxes used, in units of 10−19 erg cm−2 s−1⊓⊔′′−1, as a function of angular distance in arcsecs along the slit. We caution that the error bars shown are merely statistical noise errors and do not include the fluctuations of the sky level reflected in the difficulty of finding a "clean" patch of sky to place the background subtraction windows on. The combined profile is not very meaningful in the innermost 2′′ because of the wide variation in amplitudes, but if there were a universal fluorescent glow we would expect the outskirts of the objects to take the appearance of annuli of uniform surface brightness. In practice, such a plateau should be washed out by the distribution in sizes of the emitters. There is a hint of a flattening between 3 and 4′′, but the signal continues to dive beyond 5′′. However, at a level of ∼ 2 × 10−19erg cm−2 s−1⊓⊔′′−1, the surface brightness is rather higher by a factor of 2 − 4 than expected for fluorescent emission with the favored range of the UV background intensity. Close inspection of the frame shows that this signal appears due to some genuinely very extended individual objects, i.e., these are not artifacts of the seeing. This is consistent with the large extent of the 1.5 × 10−19erg cm−2 s−1⊓⊔′′−1 surface brightness contour, corresponding to a flux density 1.5 × 10−20erg cm−2 s−1A seen in fig. 3. – 32 – It is intriguing that the corresponding physical radius is 30 kpc, about four times larger than we had estimated based on the individually modelled surface brightness profiles above. If this were identical to the radius out to which all objects have optically thick HI, our sample of 27 emitters would project a dN/dz ∼ 1.4, and would correspond to all LLS with column densities larger than about 3 × 1018 cm−2, where the fraction of Lyα photons per ionizing photon is just flattening off to attain the maximum conversion rate (Gould & Weinberg 1996). We will defer a more detailed analysis to future study, and conclude that the large lateral median extent of our emitters is fully consistent with them being surrounded by optically thick, Lyman limit zones that radiate to within a factor two at the level predicted by Gould & Weinberg (1996), updated by latest estimates for the photoionization rate. 9. Conclusions Our longslit search for Lyα fluorescence from the intergalactic medium, taking advantage of moderate seeing periods with FORS2 at the VLT, has yielded a sample of 27 faint emitters with line fluxes of a few ×10−18 erg s−1cm−2 over a redshift range 2.66 < z < 3.75. At least a third of the sample shows emission line profiles or an association with absorp- tion systems in the nearby QSO, strongly suggesting identification with Lyα. Spectroscopic features and the absence of detected continua down to 3 − σ flux limits of ∼ 1.5 × 10−19 erg s−1cm−2 make a direct identification of the other emitters (as HI Lyα, [OII] doublet, or [OIII]/HI Balmer emission lines) difficult, but comparison with known galaxy populations and other statistical arguments indicate that the majority of emitters is likely to be Lyα at mean redshift 3.2. If this identification is correct, the emitters present a steeply rising luminosity function with a total number density more than 20 times larger than the comoving density of Lyman break galaxies (MR < 25.5) at comparable redshifts. About half of the profiles are extended, possibly owing to radiative transfer of Lyα photons from a central source, and there are candidates for both outflows and infall features. We have investigated several mechanisms for the Lyα production and find star formation to be the energetically most viable process, with a few objects being candidates for cooling radiation. The inferred low star formation rates, large line emission cross-sections, high number density, and a fitting total cross-section per unit redshift on the sky seem to provide an excel- lent match to the low luminosities, low metallicities, low dust content, and rate of incidence of damped Lyα systems, the main reservoir of neutral gas at high redshift. This suggests that our objects are the long-sought counterparts of DLAS in emission. The properties of – 33 – the objects paint the DLAS host galaxies as a population of low mass proto-galactic clumps as suggested by some of us (Haehnelt et al 1998, 2000; Rauch et al 1997b), disfavoring a model dominated by large disk galaxies (Prochaska & Wolfe 1997). Recent, apparently con- tradictory limits on spatially extended star-formation from DLAS (Wolfe & Chen 2007) and on the heating of DLAS (Wolfe, Gawiser & Prochaska 2003) are consistent with the amount of star-formation we measure if it is confined to a small unresolved region, irrespective of the fact that both the absorption and the Lyα emission cross-sections appear much larger, the latter because of the random walk of photons to the edge of optically thick gas. The physical origin and the nature of the extended optically thick gas is uncertain, but it is intriguing that quite a few objects show Lyα emission line profiles consistent with galactic outflows. Thus the absorption cross section of DLAS could be enhanced by winds, a possibility raised by Nulsen, Barcons & Fabian (1998) and discussed in the context of Lyman break galaxies by Schaye (2001). Finally, adding up the surface brightness profiles of all objects in the outer 2-6′′ we detect radiation at the > 2× 10−19 erg cm−2 s−1⊓⊔′′−1 level out to 4′′. The light level is higher by a factor 2 − 4 than the Lyα fluorescence signal expected for a UV background intensity consistent with current estimates for the photoionization rate of the intergalactic medium. The large sizes can be explained if radiative transfer of Lyα lights up the outskirts of our objects out to a radius where the conversion from UV to Lyα starts becoming inefficient. A radius of 4′′ combined with the observed comoving density of our sample can explain the rate of incidence of DLAS and LLS with column densities as low as 3× 1018cm−2, consistent with the possibility that many LLS arise in the outskirts of DLAS which in turn are to be identified with the faint emitters. With our interpretation the gas in DLAS is the reservoir from which typical L∗ galaxies must have formed in a CDM based universe. Some of our conclusions here are speculative. Further study is highly desirable, but in any case it should be obvious that long spectroscopic exposures are a promising way of discovering low-mass galaxies. Performing single field, blank sky searches for Lyα emitters for longish amounts of observing time (but not longer than regularly used by radio astronomers and now even in space based UV, optical and X-ray astronomy) can bring a whole new range of astrophysical phenomena within the range of existing ground based, optical telescopes, and obviously, could provide one of the most exciting science projects for a future generation of ultra-large telescopes. We would like to thank the following people for helpful advice and discussions during the course of this project: James Bolton, Scott Burles, Bob Carswell, Hsiao-Wen Chen, Sandro D'Odorico, Johan Fynbo, Rob Kennicutt, Juna Kollmeier, Rob Simcoe, and Neil – 34 – Trentham. We thank the Director General of ESO, Catherine Cesarsky, for keeping this project afloat, and we are grateful to the observatory staff at ESO and Gemini for performing the observations, and to the time assignment committees of both observatories. MR thanks the IoA for hospitality in spring and summer 2007. GDB and MR were supported by the National Science Foundation through grant AST 05-06845. This work was partially supported by the EC RTN network "The Physics of the Intergalactic Medium". – 35 – A. Slit Losses and Selection Effects The determination of total fluxes and surface brightnesses from long slit spectra is generally not possible for individual objects, the problem being the unkown spatial shape, and overall extent, and the random position of the slit center relative to the position of the underlying object in the plane of the sky. The spatial variation of the surface brightness along the slit is bound to generally be different from the intrinsic surface brightness dependence on radius. However, performing simulations with a given radial surface brightness distribution, a size, and radial symmetry, one can obtain an impression of the average slit losses and of the relation between the actual radial surface brightness dependence and the typical distribution of surface brightness as a function of position along the slit. The results of Monte Carlo simulations of a 2" wide, long slit, randomly positioned over a set of disks with a given size and surface brightness dependence, are shown in figures 21 and 22. The model distribution assumes the surface brightness to have a Gaussian core with a width to match the measured seeing, replaced by a power law at larger radii: and S(r) =∝ exp(−r2/(2σ2), for r < rturn, S(r) =∝ rα, f or r > rturn. (A1) (A2) For simplicity, the transition between the Gaussian and the power law at the turnover radius, rturn, was chosen such that the amplitude and slope of the two regimes were continuous at rturn, corresponding to rturn = √−ασ. Fig. 21 shows a number of radial model surface brightness distributions (smooth curves) and the predicted average observed distributions (binned curves) along a 2" wide long slit, derived from a Monte Carlo simulations with 100 realizations (= random slit positions) per model. For the input distribution, the abscissa in all panels is the radius from the center of the emitter; for the output predicted surface brightness it is the spatial coordinate along the slit (both in units of 0.252" wide pixels). Note that for non-trivial functional forms these distributions should not agree unless the slit is infinitely narrow and runs precisely radially with respect to the underlying emitter. The left column of panels shows the form of these distributions for three different exter- nal power law slopes α = -0.5 (top), -1.5 (middle), and -3.0 (bottom), representing profiles increasingly dominated by a central peak. The right column of panels shows the ratios between the same input radial surface brightness distributions and the predicted "along-the- slit" distributions. The multiple curves in each panel represent emitters with different overall radial extent, going from R = 5 to 30 pixels in steps of 5. With a pixel scale of 0.252", the – 36 – largest emitter model considered would then be 30 × 0.252" ≈ 7.5" in radius. The dotted vertical lines in the LHS panels show the location of the transition radius between central Gaussian and external power law. The RHS panels show the ratios between input radial and output slit distribution. From the RHS panels it is clear that the observed surface bright- ness along the slit is significantly distorted from the actual radial distribution. In particular, the central peak and the outer edges are suppressed in the observed profile, in both cases because they do not subtend a large area and are difficult to hit by a randomly positioned slit. Besides, the overall surface brightness is depressed typically by a factor 2 or more. Fig. 22 shows the dependence of the average total flux received through the slit as a function of the radius of the underlying emitter (again in pixels). The spectrum contains about 43 % of the flux for a source cut off radially at 5 pixels (∼ 1.25"), but only 20% for a source with an extent of 16 pixels (∼ 4"). Thus within the R=4" where there is detectable flux in some of the sources discovered here we would underestimate the total flux by up to a factor 5. When applied to a luminosity function this upward correction in flux by a factor 2 - 5 will not affect the numbers of emitter in the faint bins (say below F=10−18), because the cumulative distribution is flat (incomplete) here anyway, but it will increase each individual luminosity in the brighter bins, leading to a larger abundance of relatively bright objects, and will just shift the abundance of emitters to a flux range brighter by a factor 2 - 5. REFERENCES Adams, T. F. 1972, ApJ, 174, 439 Adelberger, K. L., Steidel, C. C., Kollmeier, J, A., Reddy, N. A., 2006, ApJ,637, 74 Ahn, S.-H., Lee, H.-W., Lee, H. M., 2003, MNRAS, 340, 863 Ahn, S.-H., 2004, ApJ, 601, L25 Beckwith, S. V. W., Stiavelli, M., Koekemoer, A. M., Caldwell, J. A. R., Ferguson, H. C., Hook, R., Lucas, R. A. Bergeron, L. E., Corbin, M., Jogee, S., Panagia, N., Robberto, M., Royle, P., Somerville, R. S., Sosey, M., 2006, AJ, 132, 1729 Binette, L., Wang, J. C. L., Zuo, L., Magris, C. G., 1993 ,AJ,105,797 Bolton, J. S., Haehnelt, M. G., Viel, M., Springel, V.,2005,MNRAS,357, 1178 Bouwens, R. J., Illingworth, G. D., Franx, M., Ford, H., 2007, ApJ, submitted, arXiv:0707.2080 Bower, R. G., Morris, S. L., Bacon, R., Wilman, R. J., Sullivan, M., Chapman, S., Davies, R. L., de Zeeuw, P. T., Emsellem, E., 2004,MNRAS, 351, 63 – 37 – Brocklehurst, M., 1971,MNRAS, 153, 471 Bunker, A. J., Marleau, F. R., Graham, J. R., 1998, AJ, 116, 2086 Bunker, A. J., Marleau, F. R., Graham, J. R., 1999, in: proceedings of the "Galaxies in the Young Universe II" workshop, Ringberg Castle 2-6 August 1999, Lecture Notes in Physics, eds. H. Hippelein & K. Meisenheimer Bunker, A. J., Warren, S. J., Clements, D. L., Williger, G. M., Hewett, P. C., 1999, MNRAS, 309, 875 Bunker, A., Smith, J., Spinrad, H., Stern, D., Warren, S.,2003, Ap&SS, 284, 357 Bunker, A. J., Stanway, E. R., Ellis, R. S., McMahon, R. G., 2004, MNRAS, 355, 374 Cantalupo, S., Lilly, S. J., Porciani, C., 2005, ApJ, 628, 61 Cantalupo, S., Lilly, S. J., Porciani, C., 2007,ApJ,657,135 Christensen, L., Wisotzki, L., Roth, M. M., Sanchez, S. F., Kelz, A., Jahnke, K., 2007, A&A, 468, 587 Chen, H.-W., Kennicutt, R.C., Rauch, M., 2005,ApJ, 620, 703 Chun, M. R., Gharanfoli, S., Kulkarni, V. P., Takamiya, M., 2006, AJ, 131, 686 Cowie, L. L., Hu, E. M., 1998, AJ,115, 1319 Dey, A., Bian, C., Soifer, B.T., Brand, K., Brown, M.J.I., Chaffee, F.H., LeFloc'h. E., Hill, G., Houck, J.R., Jannuzi, B.T., Rieke, M., Weedman, D., Brodwin, M., Eisenhardt, P., 2005, ApJ, 626, 654 Dijkstra M., Haiman, Z., Spaans, M. , 2006, ApJ, 649, 14 Dijkstra M., Haiman, Z., Spaans, M. , 2006, ApJ, 649, 37 Erb, D. K., Steidel, C. C., Shapley, A. E., Pettini, M., Reddy, N. A., Adelberger, K. L., 2006, ApJ, 128 Fardal, M. A., Katz, N., Gardner, J. P., Hernquist, L., Weinberg, D. H., Dav, R., 2001, ApJ, 562, 605 Francis P. J., et al., 2001, ApJ, 554, 1001 Francis, P. J., Bland-Hawthorn, J., 2004, MNRAS, 353, 301 Francis, P. J., McDonnell, S., 2006,MNRAS, 370, 1372 Franx, M., Illingworth, G. D., Kelson, D. D., van Dokkum, P. G.; Tran, K.-V. 1997, ApJ,486, 75 Fujita, S. S., Ajiki, M., Shioya, Y., Nagao, T., Murayama, T., Taniguchi, Y., Okamura, S., Ouchi, M., Shimasaku, K., Doi, M., and 18 other authors, 2003,AJ, 125, 13 – 38 – Furlanetto, S. R., Schaye, J., Springel, V., Hernquist, L., 2005, ApJ, 622, 7 Fynbo, J. U., Møller, P., Warren, S. J., 1999, MNRAS, 305, 849 Fynbo, J. P. U., Ledoux, C., Mller, P., Thomsen, B., Burud, I., 2003 A&A, 407,147 Gerhard, O., Arnaboldi, M., Freeman, K. C., Okamura, S., Kashikawa, N., Yasuda, N.,2005, ApJ, 621, 93 Gerhard, O., Arnaboldi, M., Freeman, K. C., Okamura, S., Kashikawa, N., Yasuda, N., 2007, A&A,468, 815 Gould, A., Weinberg, D. H., 1996, ApJ, 468, 462 Gronwall, C., Ciardullo, R., Hickey, T., Gawiser, E., Feldmeier, J. J., van Dokkum, P. G., Urry, C. M., Herrera, D., Lehmer, B. D., Infante, L., and 6 coauthors, 2007, ApJ, in press, arXiv:0705.3917 Haardt, F., Madau, P., 1996, ApJ, 461, 20 Haehnelt, M. G., Steinmetz, M., Rauch, M., 1998,ApJ, 495, 647 Haehnelt, M. G., Steinmetz, M., Rauch, M., 2000, ApJ, 534, 594 Haiman, Z., Rees, M. J., 2001, ApJ, 556, 87 Haiman, Z., Spaans, M., Quataert, E., 2000, ApJ, 537, 5 Hall, P. B., Osmer, P. S. Green, R. F.; Porter, A. C.; Warren, S. J., 1996, ApJ, 462, 614 Hennawi, J. private communication, 2007 Hogan, C. J., Weymann, R. J., 1987, MNRAS, 225, 1 Hogg, D. W., Cohen, J. G., Blandford, R., Pahre, M. A., 1998, ApJ, 504, 622 Hu, E. M., Cowie, L. L., McMahon, R. G., 1998,ApJ, 502, 99 Johansson, P. H.; Efstathiou, G., 2006, MNRAS, 371, 1519 Kauffmann, G., 1996, MNRAS, 281, 475 Keel, W. C., Cohen, S. H., Windhorst, R. A., Waddington, I., 1999, AJ, 118, 2547 Kelson, D. D.,2003, PASP, 115, 688 Kennicutt, R.C., 1998, ApJ,498, 541 Kudritzki, R.-P., Mendez, R. H., Feldmeier, J. J., Ciardullo, R., Jacoby, G. H., Freeman, K. C., Arnaboldi, M., Capaccioli, M., Gerhard, O., Ford, H. C., 2000,ApJ, 536, 19 Kulkarni, V. P., Hill, J. M., Schneider, G. Weymann, R. J., Storrie-Lombardi, L. J., Rieke, M. J., Thompson, R. I., Jannuzi, B. T., 2000, ApJ, 536, 36 – 39 – Kulkarni V. P., Hill, J. M., Schneider, G., Weymann, R. J., Storrie-Lombardi, L. J., Rieke, M. J., Thompson, R. I., Jannuzi, B. T., 2001, ApJ, 551, 37 LeBrun v., Bergeron, J., Boiss´e, P., Deharveng, J.M., 1997, A&A, 321, 733 Le Delliou, M., Lacey, C., Baugh, C.M., Guiderdoni, B., Bacon, R., Courtois, H., Sousbie, T., and Morris, S.L., 2005, MNRAS, 357, L11 Le Delliou, M., Lacey, C.G., Baugh, C.M., and Morris, S.L., 2006, MNRAS, 365, 712 Ledoux C., Petitjean P., Bergeron, J., Wampler E. J., Srianand, R., 1998, A&A, 337, 51 Lequeux, J., Kunth, J.M., Mas-Hesse, J.M., Sargent, W.L.W., 1995, A&A, 301, 18 Lowenthal, J., D. Hogan, C. J., Leach, R. W., Schmidt, G. D., Foltz, C. B., 1990,ApJ,357,3 Madau, P., Pozzetti, L., Dickinson, M., 1998,ApJ, 498, 106 Mas-Hesse, J.M., Kunth, D., Tenorio-Tagle, G., Leitherer, C., Terlevich, R.J., Terlevich, E., 2003, ApJ, 598, 858 Matsuda, Y., Yamada, T., Hayashino, T., Yamauchi, R., Nakamura, Y., 2006,ApJ640, 123 Møller, P., Warren, S. J., Fall, S. M., Fynbo, J. U., Jakobsen, P., 2002, ApJ574, 51 Mo, H. J., White, S. D. M., 2002, MNRAS, 336, 112 Murphy, M.T., Liske, J., 2004,MNRAS, 354, 31 Nagamine K., Wolfe A.M., Hernquist L. , Springel, V., 2007, ApJ, 660, 945 Neufeld, D.A., 1990, ApJ, 350, 216 Nilsson, K. K., Fynbo, J. P. U., Møller, P., Sommer-Larsen, J., Ledoux, C.,2006,A&A,452,23 Nulsen, P.E.J., Barcons, X., Fabian, A.C., 1998, MNRAS, 301, 168 Ouchi, M., Shimasaku, K., Akiyama, M., Simpson, C., Saito, T., Ueda, Y., Furusawa, H., Sekiguchi, K., Yamada, T., Kodama, T., and 6 coauthors, submitted to ApJ, arXiv:0707.3161 Peroux, C., McMahon, R. G., Storrie-Lombardi, L. J., Irwin, M. J., 2003, MNRAS. 346, 1103 Peroux C., Dessauges-Zavadsky M., D'Odorico S., Kim, T.-S.,, McMahon, R. G., 2005,MN- RAS,363, 479 Prochaska, J. X., Wolfe, A. M., 1997, ApJ, 487, 73 Rao S.M., Turnshek D.A., Nestor D.B., 2006, ApJ, 636, 610 Rauch, M., Haehnelt, M. G., Steinmetz, M. 1997a, ApJ, 481, 601 – 40 – Rauch, M., Miralda-Escude J., Sargent, W. L. W., Barlow, T. A.; Weinberg, D. H., Hern- quist, L., Katz, N., Cen, R., Ostriker J. P., 1997b, ApJ, 489, 7 Schaye, J., 2001, ApJ, 559, 1 Scott, J. E., Kriss, G. A., Brotherton, M., Green, R. F., Hutchings, J., Shull, J. M., Zheng, W., 2004, ApJ, 615, 135 Shapley, A. E., Steidel, C. C., Pettini, M., Adelberger, K. L., 2003, ApJ, 588, 65 Shapley, A. E., Steidel, C. C., Pettini, M., Adelberger, K. L., Erb, D. K., 2006, ApJ, 651, 688 Simcoe, R. A., Sargent, W. L. W., Rauch, M., 2004,ApJ,606, 92 Smith, D. J. B., Jarvis, M. J., 2007, MNRAS,378, 49 Stark, D. P., Ellis, R. S., Richard, J., Kneib, J.-P., Smith, G. P., Santos, M. R., 2007, ApJ, 663, 10 Steidel, C. C., 1990, ApJS, 74, 37 Steidel, C. C., Hamilton, D., 1993, AJ, 105,2017 Steidel, C. C., Adelberger, K. L., Giavalisco, M., Dickinson, M., Pettini, M., 1999,ApJ, 519, 1 Steidel, C. C., Adelberger, K. L., Shapley, A. E., Pettini, M., Dickinson, M., Giavalisco, M., 2000, ApJ, 532, 170 Stiavelli, M., Scarlata, C., Panagia, N., Treu, T., Bertin, G., Bertola, F., 2001,ApJ, 561, 37 Tasitsiomi, A., 2006, ApJ, 645, 792 Tapken, C., Appenzeller, I., Noll, S., Richling, S., Heidt, J., Meinkohn, E., Mehlert, D., 2007, A&A, 467, 63 Telfer, R. C., Zheng, W., Kriss, G. A., Davidsen, A. F., 2002 ApJ, 565, 773 Tenorio-Tagle, G., Silich, S. A., Kunth, D., Terlevich, E., Terlevich, R., 1999, MNRAS, 309, 332 Trentham, N., Sampson, L., Banerji, M., 2005, MNRAS, 357,783 van Breukelen, C., Jarvis, M.J., Venemans, B.P., 2005, MNRAS, 359, 895 Wang, L., Li, C., Kauffmann, G., De Lucia, G., 2007, MNRAS, 377, 1419 Warren, S. J., Møller, P., Fall, S. M.; Jakobsen, P., 2001 MNRAS, 326, 759 Weatherley, S. J., Warren, S. J., Møller, P., Fall, S. M., Fynbo, J. U., Croom, S. M., 2005, MNRAS, 358, 985 – 41 – Weidinger, M. Møller, P., Fynbo, J. P. U., 2004, Nature, 430, 999 Wild, V., Hewett, P.C., Pettini, M., 2007, MNRAS, 374, 292 Weidinger, M., Møller, P., Fynbo, J. P. U., Thomsen, B., 2005, A&A, 436,825 Wolfe, A. M., Lanzetta, K. M., Foltz, C. B., Chaffee, F. H., 1995, ApJ,454,698 Wolfe, A. M., Gawiser, E, Prochaska, J. X., 2003, ApJ, 593, 235 Wolfe, A. M., Chen, H.-W., 2006, ApJ, 652, 981 Zheng, Z., Miralda-Escud´e, J., 2002, ApJ, 578, 33 This preprint was prepared with the AAS LATEX macros v5.2. – 42 – Fig. 1.- Two-dimensional spectrum obtained in 92 hours of exposure time., showing the line emitter candidates for HI Lyα (boxes). The dispersion direction is horizontal, with blue to the left and red to the right; the spatial direction along the slit is vertical. The QSO spectrum (multiple absorption lines) is visible close to the center of the image. The numbers refer to the column entry 'ID' in table 2. – 43 – Fig. 2.- Spectra of the emission line regions of five foreground line emitting galaxies identified from their [OII] doublet (objects 7a, 7b, 31) or Balmer emission (Hβ for object 8, Hγ for object 22) features. The coordinates are in pixel units (0.252′′ × 0.67A). The sections of the spectra shown here are 15.12" wide in the spatial direction and about 2266 kms−1 long in the spectral direction (i.e., horizontally). The numbers refer to the column entry 'ID' in table 1. – 44 – Fig. 3.- short spectra for the single line emitters. The coordinates are in pixel units (0.252′′ × 0.67A). The sections of the spectra shown here are 15.12" or 116 proper kpc wide in the spatial direction and about 2266 kms−1 long in the spectral direction (i.e., horizontally). The spectra have been heavily smoothed with a 7x7 pixel boxcar filter. The areas within the light grey contours (turqoise in the color version) have a flux density greater than approximately 1.5× 10−20 erg cm−2 s−1 A. The numbers refer to the column entry 'ID' in table 2. The spectra are grouped together such that the first 12 of them (top box) appear to have a single central peak; the next six (IDs 39, 27, 3, 23, 28, and 29, second box from top) show a clearly asymmetric red peak with a much weaker blue counter-peak; the following three (third box to the left) either have a stronger blue than red peak (ID 15) or emission features blueward of an absorption line (36, 37); the remaining six are unclassifiable, sometimes amorphous objects. – 45 – Fig. 4.- two-dimensional spectra, extracted 1-d spectra (flux density in erg cm−2 s−1 A−1), and spatial surface brightness cross-sections (in erg cm−2 s−1 ⊓⊔′′−1) along the slit. The spectra are about 25.6 A long in the horizontal direction and 7.56" vertically, and are smoothed with a 3x3 pixel filter. The solid, dashed, and dotted lines show the actual data, a Gaussian PSF normalized to the surface brightness in the central 2 pixels, and a fitted model consisting of that central Gaussian and a power law continuation further out, respectively. – 46 – Fig. 4.- continued. – 47 – Fig. 4.- continued. – 48 – s t c e b o j f o r e b m u n 10 8 6 4 2 0 4600 4800 5000 5200 5400 5600 5800 observed wavelength [A] Fig. 5.- Distribution of the emitters in observed wavelength. The solid line is the sensitivity of the instrument, in arbitrary units. The actual detection threshold probably drops faster toward the edges because of illumination, dithering losses, and detector artifacts. – 49 – 6 4 2 0 0 1 radius along the slit [arcsec] for SB=10-19 2 3 4 Fig. 6.- distribution of the projected radius along the slit, i.e., the distance of the 10−19 surface brightness contour (based on the model profile) from the center of the emitter. The four cases where the contour extends beyond our fitting range are collected in the 3.6′′bin, but the contour may reach considerably larger distances than that. – 50 – 2.0 1.5 1.0 i ) ] 2 - n m c r a [ ) Ω ∂ z ∂ ( / ) α y L F > ( N ∂ ( 2 g o l -19.0 -18.5 -18.0 -17.5 log (FLyα [erg cm-2 s-1]) -17.0 Fig. 7.- Number of emitters per unit redshift and square arcminute with a line flux exceeding FLyα. – 51 – 8 6 4 2 0 0 50 100 LLyα[1040 h70 -2 erg s-1] 150 200 Fig. 8.- Frequency distribution of luminosities LLyα. – 52 – -1.0 -1.5 -2.0 -2.5 -3.0 ) ] 3 - c p M 0 3 7 h [ c V ∂ / ) α y L L > ( N ∂ ( g o l 0.5 1.0 log (LLyα [1040 h70 -2 erg s-1]) 1.5 2.0 Fig. 9.- Comoving density of emitters with a luminosity exceeding LLyα. – 53 – ) ] 3 - c p M 0 3 7 h [ c V ∂ / ) ] I I [ O L > ( N ∂ ( g o l -0.5 -1.0 -1.5 -1.5 -1.0 -0.5 log (L[OII] [1040 h70 -2 erg s-1]) Fig. 10.- Comoving density of emitters under the assumption that they are [OII] 3728 A, with a luminosity exceeding L[OII]. – 54 – 1.0 0.5 0.0 ) ] 3 - c p M 0 3 7 h [ c V ∂ / ) ] I I I [ O L > ( N ∂ ( g o l -4.0 -3.5 log (L[OIII] [1040 h70 -3.0 -2.5 -2 erg s-1]) -2.0 Fig. 11.- Comoving density of emitters under the assumption that they are [OIII] 5007 A, with a luminosity exceeding L[OIII]. – 55 – i ) z d ( / ) s u d a r > ( N d ( g o l 1.0 0.8 0.6 0.4 0.2 0.0 0 5 10 20 radius [phys. kpc]) 15 25 30 Fig. 12.- Contribution of objects of different sizes to the rate of incidence per unit redshift, dN/dz, for HI (solid line), [OII] (dashed line), and [OIII] (dotted line). The short vertical lines riding on top of the curves indicate the spatial resolution limit along the slit. – 56 – -1 -2 -3 -4 -5 ] 3 - c p M 0 3 7 h [ ) V d / ) L > ( N d ( g o l 0 1 log( LLyα [1040h70 3 -2 erg s-1]) 2 4 Fig. 13.- Observed cumulative luminosity function at z=3.1 from Ouchi et al (2007; dashed line), z=2.9 from van Breukelen et al (2005; dash-triple dotted line), predictions from LeDelliou et al (2005, 2006; dash-dotted line) and our sample (solid line, with the dotted lines representing the 1 − σ error contours). There is almost continuity in amplitude and slope in the overlap region with the brighter observational data. The van Breukelen et al function with its adopted α = −1.6 slope if continued beyond its measured range is never more than 0.2 dex below our curve. The theoretical curve from LeDelliou et al for a constant escape fraction is shallower than all the observed distributions at all luminosities, but the gap between it and our function steepens toward fainter luminosities. Our observed distribution starts to flatten near 3× 1041erg s−1, well above the detection threshold, so the turnover may be intrinsic. – 57 – -1.5 -2.0 -2.5 ) ] 1 - z H 1 - s g r e 0 2 7 - h 0 3 0 1 [ V L ( g o l -3.0 0.5 1/10 Mo yr-1 4/10 Mo yr-1 1.0 log(LLyα [1040 h70 1.5 -2 erg s-1]) 2.0 Fig. 14.- V-band luminosity density versus Lyα line luminosities, under the assumption that all sources are high redshift Lyα. The arrows are upper limits, i.e., detected line emitters without V band counterparts at a 3−σ significance level (in a 2′′×2′′ aperture). Of the three positive detections 39, 9, and 33, the latter is somewhat off center and may be a low z source or a chance coincidence of a high redshift emitter with an lower redshift continuum source. Three objects, 1, 3 and 39 are not covered by the V band image. The spectrum of 39, however, shows a continuum consistent with the expected Lyα forest decrement. We have given here its 1500A rest frame luminosity measured from the spectrum instead of the V band luminosity. The lower, dashed, diagonal line is the expected locus of Lyα emitters assuming that both, UV luminosity and Lyα flux, were produced by star formation only (see text). This line corresponds to a rest frame equivalent width of 68 A and would intercept the y-axis at -4.14. The higher dashed line with the same slope delineates EW=20A. The vertical dotted lines in Lyα emitters indicate star formation rates of 1/10 (left) and 4/10 (right) M⊙ yr−1. – 58 – ] 3 - c p M 0 3 7 h [ ) V d / ) M < ( N d ( g o l -1 -2 -3 -4 -5 -6 -22 -20 MUV,AB -18 -16 Fig. 15.- Cumulative UV continuum luminosity functions of Steidel et al (1999; asterisks), the Hubble Ultra Deep Field (Bouwens et al 2007; dashed line), and the cumulative distribution of our survey (solid line; dotted lines are ±1σ errors). The diamond symbol shows object # 39. The emitters are entered with a continuum magnitude predicted by their Lyα line flux as we have only upper limits on the continuum - see text. The number density of our emitters closely corresponds to the number density of Bouwens et al (2007). The absence of objects brighter than -21 is consistent with our small survey volume. – 59 – 12 16 QSO Fig. 16.- The elliptical 'zone of influence' near the QSO where the ionizing radiation would be sufficient to cause fluorescence at the 10−18erg cm−2 s−1 arcsec−2 surface brightness level. – 60 – l a h p a x e d n i w a l r e w o p 0 -1 -2 -3 -4 -5 -6 0 500 1000 1500 velocity width [km s-1] 2000 Fig. 17.- Power law slope for the Gaussian + power law fit to the spatial surface brightness profile, versus Gaussian velocity width of the emission line. There is no significant correlation between the two, but the objects with the largest velocity widths also seem to have small power law indices, i.e., the slowest radial decline in surface brightness . – 61 – 36 37 QSO A 15 B C D Fig. 18.- The immediate surroundings of the smudges # 36, 37, and 15. The labels A,B,C, and D refer to the strong absorption lines in the QSO spectrum, partly lining up with the emitters (see text). The featureless continuum object just above the QSO is a low redshift object, the fainter continuum further up appears to show absorption by the emitting smudges 36 and 37. – 62 – i ) z d ( / ) s u d a r > ( N d ( g o l 0.8 0.6 0.4 0.2 0.0 0 10 20 radius [phys. kpc]) 30 Fig. 19.- Contribution of objects of different sizes to the rate of incidence per unit redshift, dN/dz, for HI with (dotted line) and without (solid line) correction for the extended sizes and our underestimating the radius. The short vertical line riding on top of the uncorrected curve indicates the spatial resolution limit along the slit. – 63 – ] 9 1 - 0 1 [ B S n a d e m i , n a e m 8 6 4 2 0 1 2 3 4 radius along slit [arcsec] 5 6 7 Fig. 20.- Mean (points with error bars) and median (open squares) surface brightness measure- ments in units of 10−19 erg cm−2s−1⊓⊔′′−1 for the combined surface brightness profiles, as a function of angular distance in arcsecs from the center of emission along the slit. The dotted lines give the range of the expected surface brightness based on the Bolton et al (2005) photoionization rate of the z ∼ 3 IGM. The upper dotted line is for a QSO type UV spectrum, the bottom line for a spectrum where 50% of the flux is contributed by galaxies. The dashed line is the 4σ surface brightness detection threshold for individual objects. – 64 – Fig. 21.- Simulation of slit losses. Left hand side column: radial surface brightness profile of the input model (smooth curve) and 'observed' average profiles (binned curves) along the slit after passage through a finite slit (2′′or about 8 pixels wide). RHS column: ratios between the input and observed profile as a function of distance along the slit. The three rows of panels in either column show models with external slope -0.5 (top row),-1.5 (middle row), and -3.0 (bottom row). The different output profiles within each panel arise from emitters with different overall radial extent. Dashed vertical lines show the position of turnover between Gaussian core and power law wings. – 65 – Fig. 22.- Average fraction of total flux passing through the 2"(8 pixel) wide slit, as a function of total radial extent of the emitter. Table 1. Foreground Emission Line Objects (1) # 1 2 3 4 5 (2) ID 7a 7b 8 22 31 (3) z 0.3928 0.4336 0.0458 0.1980 0.4019 (4) (5) (6) F [10−18]a Smax [10−18] Source of Identification 12.45 ± 0.29 27.53 ± 0.33 2.60 ± 0.28 18.49 ± 0.31 26.55 ± 0.32 [OII] doublet 5.06 ± 0.22 [OII] doublet 8.99 ± 0.23 1.10 ± 0.21 Hβ and Mgb triplet 8.30 ± 0.23 Balmer lines, [OII] doublet 9.24 ± 0.23 [OII] doublet atotal flux for both lines where doublet – 6 6 – Table 2. Properties of Single-Line Emitters (1) # (2) ID (3) z (4) (5) (6) (7) (8) F [10−18] flux ratio Smax [10−18] vF W HM A [10−18] 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 1 3 4 6 9 10 12 14 15 16 17 18 19 20 21 23 24 25 26 27 28 29 30 33 36 37 38 39 3.1801 3.1916 3.1797 3.3362 3.2378 3.4833 3.3300 3.3221 2.7682 3.3189 3.6954 3.4373 3.0595 3.4023 3.0809 2.9075 3.0593 3.0655 3.0692 3.2617 3.0732 3.1819 3.2715 3.2646 2.7483 2.7713 3.0322 2.8285 2.32 ± 0.57 12.77 ± 0.34 3.38 ± 0.34 6.04 ± 0.35 2.89 ± 0.35 1.91 ± 0.38 3.36 ± 0.35 2.67 ± 0.41 3.06 ± 0.37 4.12 ± 0.45 1.14 ± 0.37 1.53 ± 0.44 1.79 ± 0.37 2.77 ± 0.36 15.32 ± 0.37 1.48 ± 0.35 0.70 ± 0.36 -0.04 ± 0.37 3.51 ± 0.34 2.90 ± 0.36 3.96 ± 0.34 3.05 ± 0.35 3.50 ± 0.34 3.46 ± 0.52 3.27 ± 0.57 3.43 ± 0.70 2.28 ± 0.43 0.53 ± 0.13 0.73 ± 0.02 0.47 ± 0.05 0.67 ± 0.04 0.79 ± 0.10 0.56 ± 0.11 1.29 ± 0.13 0.38 ± 0.06 0.59 ± 0.07 0.62 ± 0.07 0.50 ± 0.16 3.27 ± 0.94 0.37 ± 0.08 0.72 ± 0.09 0.65 ± 0.02 0.69 ± 0.16 0.68 ± 0.35 -0.22 ± 2.02 0.78 ± 0.08 0.45 ± 0.06 0.48 ± 0.04 1.56 ± 0.18 1.00 ± 0.10 0.81 ± 0.12 1.14 ± 0.20 2.69 ± 0.55 0.41 ± 0.08 aEXT = 'extended' bCD = 'centrally dominated' cPS = 'point source' 0.88 ± 0.4 4.81 ± 0.24 1.14 ± 0.24 2.09 ± 0.24 0.81 ± 0.24 0.67 ± 0.26 1.30 ± 0.24 1.02 ± 0.29 1.08 ± 0.28 1.40 ± 0.32 0.48 ± 0.27 0.47 ± 0.30 0.68 ± 0.26 1.29 ± 0.25 5.30 ± 0.26 0.58 ± 0.25 0.28 ± 0.25 0.19 ± 0.25 1.23 ± 0.24 1.07 ± 0.25 1.36 ± 0.24 1.27 ± 0.25 1.88 ± 0.24 0.90 ± 0.31 0.73 ± 0.29 1.44 ± 0.51 0.75 ± 0.33 352.2 382.1 431.8 289.8 288.3 312.4 296.7 883.4 528.3 493.9 289.3 158.8 514.3 228.1 444.4 228.8 357.1 238.9 249.2 885.7 279.5 313.1 322.2 1073.7 725.0 2151.1 275.6 1.04 4.81 1.28 2.26 0.92 0.81 1.15 1.04 1.48 1.48 0.24 0.13 0.69 1.64 5.45 0.53 0.11 fit 1.32 1.09 1.57 0.88 1.51 1.08 1.02 0.79 0.86 (9) yt 3.0 3.9 2.9 3.9 2.2 6.5 6.5 2.4 4.1 2.7 2.0 3.1 2.1 8.5 2.5 2.5 4.2 failed 4.7 2.2 3.3 4.4 4.0 2.7 4.9 5.2 2.2 (10) α -2.12 -2.07 -0.50 -2.12 -3.78 -3.87 -3.75 -0.91 -2.20 -1.68 -0.54 -2.20 -0.99 -5.37 -1.46 -1.78 -4.05 - -2.49 -1.09 -0.34 -2.38 -2.16 -1.62 -2.76 -2.70 -1.07 (11) Comments somewhat EXTa , CDb PSc ; red-dominated em. w. faint blue peak EXT CD CD, coincidence with unrelated (?) continuum object EXT, "plug-shaped" em. PS QSO centered on broad em. line EXT "ring" with blue-dominated double comp. CD; broad em. line EXT; broad em. line EXT, amorphous CD, somewhat EXT EXT, amorphous PS; narrow line CD, ring; weak blue, strong red double comp. PS, faint PS, faint PS, faint PS; narrow line – 6 7 – CD, somewhat EXT; weak blue, strong red double comp. CD; weak blue, strong narrow red double comp. CD PS; coincidence with unrelated lower z continuum obj. very EXT, amorphous smudge, nearby QSO and gal. abs. EXT, assym.; em. ('trapdoor') blueward of abs. very broad em. on top of fuzzy continuum CD, narrow PCygni em. line w. continuum
astro-ph/0106224
1
0106
2001-06-13T08:27:28
On the formation of oxygen-neon white dwarfs in close binary systems
[ "astro-ph" ]
The evolution of a star of initial mass 10 $M_{\odot}$, and metallicity $Z = 0.02$ in a Close Binary System (CBS) is followed from its main sequence until an ONe degenerate remnant forms. Restrictions have been made on the characteristics of the companion as well as on the initial orbital parameters in order to avoid the occurrence of reversal mass transfer before carbon is ignited in the core. The system undergoes three mass loss episodes. The first and second ones are a consequence of a case B Roche lobe overflow. During the third mass loss episode stellar winds may play a role comparable to, or even more important than Roche lobe overflow. In this paper, we extend the previously existing calculations of stars of intermediate mass belonging to close binary systems by following carefully the carbon burning phase of the primary component. We also propose different possible outcomes for our scenario and discuss the relevance of our findings. In particular, our main result is that the resulting white dwarf component of mass $1.1 M_\odot$ more likely has a core composed of oxygen and neon, surrounded by a mantle of carbon-oxygen rich material. The average abundances of the oxygen-neon rich core are $X({\rm O}^{16})=0.55$, $X({\rm Ne}^{20})=0.28$, $X({\rm Na}^{23})=0.06$ and $X({\rm Mg}^{24})=0.05$. This result has important consequences for the Accretion Induced Collapse scenario. The average abundances of the carbon-oxygen rich mantle are $X({\rm O}^{16})=0.55$, and $X({\rm C}^{12})=0.43$. The existence of this mantle could also play a significant role in our understanding of cataclysmic variables.
astro-ph
astro-ph
Abstract. The evolution of a star of initial mass 10 M⊙, and metallicity Z = 0.02 in a Close Binary System (CBS) is followed from its main sequence until an ONe degener- ate remnant forms. Restrictions have been made on the characteristics of the companion as well as on the initial orbital parameters in order to avoid the occurrence of re- versal mass transfer before carbon is ignited in the core. The system undergoes three mass loss episodes. The first and second ones are a consequence of a case B Roche lobe overflow. During the third mass loss episode stellar winds may play a role comparable to, or even more important than Roche lobe overflow. In this paper, we extend the pre- viously existing calculations of stars of intermediate mass belonging to close binary systems by following carefully the carbon burning phase of the primary component. We also propose different possible outcomes for our scenario and discuss the relevance of our findings. In particular, our main result is that the resulting white dwarf component of mass 1.1 M⊙ more likely has a core composed of oxygen and neon, surrounded by a mantle of carbon -- oxygen rich material. The average abundances of the oxygen -- neon rich core are X(O16) = 0.55, X(Ne20) = 0.28, X(Na23) = 0.06 and X(Mg24) = 0.05. This result has important conse- quences for the Accretion Induced Collapse scenario. The average abundances of the carbon -- oxygen rich mantle are X(O16) = 0.55, and X(C12) = 0.43. The existence of this mantle could also play a significant role in our understand- ing of cataclysmic variables. Key words: stars: evolution -- stars: binaries: general -- stars: white dwarfs 1 0 0 2 n u J 3 1 1 v 4 2 2 6 0 1 0 / h p - o r t s a : v i X r a A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: missing; you have not inserted them ASTRONOMY AND ASTROPHYSICS On the formation of oxygen -- neon white dwarfs in close binary systems Pilar Gil -- Pons1 and Enrique Garc´ıa -- Berro1,2 1Departament de F´ısica Aplicada, Universitat Polit`ecnica de Catalunya, c/Jordi Girona s/n, M´odul B-4, Campus Nord, 08034 Barcelona, Spain, (e-mail: pilar, [email protected]) 2Institute for Space Studies of Catalonia, c/Gran Capit´a 2 -- 4, Edif. Nexus 104, 08034 Barcelona, Spain Received 22 February 2001 / Accepted 25 May 2001 1. Introduction Intermediate mass close binaries are defined as those sys- tems in which the primary component develops a degen- erate carbon -- oxygen core, after burning central helium in non -- degenerate conditions. From the orbital parameters in these systems, we see that periods are small enough that the possibility of mass transfer due to Roche Lobe overflow is enabled. The evolution of low -- to intermediate -- mass stars belonging to close binary systems has been widely studied so far and, even though many questions still re- main unsolved, important contributions have already been made on this subject. One of these questions concerns the evolution of heavy -- weight intermediate mass stars (that is, primary stars with masses between ∼ 8 and 11 M⊙) belonging to close binary systems. For the case of isolated stars, this range of stellar masses corresponds to stars for which, after exhaustion of central helium, carbon is ig- nited under conditions of partial degeneracy. Ultimately, these stars become Super -- AGB stars with ONe cores (Ri- tossa, Garc´ıa -- Berro & Iben 1996). For the case of stars within this mass range belonging to binary systems, very few comments can be made, either because most of the cal- culations do not follow the evolution through the carbon burning phase or, simply, because the existing calculations focus mostly on a lower segment of masses. For instance, Whyte & Eggleton (1980) studied the evolution of stars of up to 3 M⊙ belonging to semide- tached systems. These authors later extended their work to more general scenarios in which accretion and mass transfer between low mass contact binaries were included (Whyte & Eggleton 1985). Van der Linden (1987) also per- formed conservative Case B evolutionary calculations for several masses of the primary, ranging from 3 M⊙ to 12 M⊙, but the evolution through the carbon burning phase was not followed. Besides the work they have performed in the field of massive binaries, de Loore & Vanbeveren (1992, 1994, and references therein) have also focused on the evolution of intermediate mass close binary systems (de Loore & Vanbeveren 1995). However, they were only able to follow the evolution of the primary star until the exhaustion of the helium in the core. Their calculations in- cluded both non -- conservative (de Loore & de Greve 1992) and conservative mass transfer (de Loore & Vanbeveren 1995). Very recently, Nelson & Eggleton (2001) have per- formed a very comprehensive work on intermediate mass close binary systems, exploring 5 500 evolutionary tracks of mostly Case A conservative mass transfer systems. The upper mass limit in this case was ∼ 50 M⊙ but, again, in most of the cases the evolution during the carbon burning phase was not followed or it was aborted earlier (when the carbon luminosity exceeded 1 L⊙). In another recent work Han, Tout & Eggleton (2000) determined the final mass relation for binary systems with the mass of the com- ponents ranging between 3 and 8 M⊙, starting mass loss at different times of the Hertzsprung -- Russell gap. How- ever, in these studies the assumption of conservative mass transfer was adopted and justified by the conditions of the case they consider which, in spite of corresponding to an important portion of the real cases, cannot account for all of them. Iben (1985, 1991) has extensively reviewed the physi- cal mechanisms relevant to binary systems and has thor- oughly discussed the evolution of intermediate mass close binary systems, offering an excellent overview of their evo- lution until very late stages, proposing several different scenarios and providing their probabilities of occurrence. Also, Iben & Tutukov (1984) have proposed different evo- lutionary scenarios for heavy -- weight intermediate mass close binary systems as progenitors of SNe Ia. In spite of the fact that this mass interval contains a good frac- tion of the stars which are massive enough to ignite car- bon in a non -- explosive way, the evolution of these sys- tems has been much neglected until very recently. The pioneering works of Miyaji et al. (1980) and Woosley, Weaver & Taam (1980) lead to the conclusion that the stars of this mass interval belonging to close binary sys- tems would lose most of their mass and, moreover, would develop electron -- degenerate ONe cores after the carbon burning phase. In a second phase, due to accretion from the secondary, the central density would increase until the threshold for electron capture on 24Mg and 24Na would be Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs 3 reached first, and on 20Ne and 20F later. The process of electron capture on these nuclei would finally trigger the explosive ignition of neon and oxygen at very high den- sities (ρc >∼ 2 × 1010 g cm−3). At these densities, fast electron captures on the incinerated material would bring the Chandrasekhar mass below the actual mass of the ONe core and induce gravitational collapse to neutron star dimensions. Although there is a general agreement that electron -- capture induced collapse could be successful, a major drawback of this scenario is that no detailed pre -- collapse models existed in the literature. For instance, in most of the calculations, the evolution during the mass loss phase was not followed in full detail (Nomoto 1984) or the evolution during the carbon burning phase was oversimpli- fied by introducing the so -- called steady burning approx- imation (Saio & Nomoto 1998). Therefore, the depicted evolutionary scenario could be substantially modified due to the presence of these approximations. Very recently, the evolution leading to the formation of white dwarfs with ONe cores in close binary systems is be- ing reinvestigated. For instance, Dom´ınguez, Tornamb´e & Isern (1994) have studied the formation of an ONe white dwarf through mass transfer in a close binary system. On the other hand, in a series of recent papers (Garc´ıa -- Berro & Iben 1994; Ritossa, Garc´ıa -- Berro & Iben 1996; Garc´ıa -- Berro, Ritossa & Iben 1997; Iben, Ritossa & Garc´ıa -- Berro 1997; Ritossa, Garc´ıa -- Berro & Iben 1999) the evolution of isolated heavy -- weight intermediate mass stars has also been carefully followed from the main sequence phase up to exhaustion of carbon in the core. Perhaps one of the most important conclusions of these papers is that iso- lated white dwarfs with masses MWD >∼ 1.0 M⊙ would most likely have an ONe core surrounded by a mantle of carbon -- oxygen rich material. This bears important conse- quences for the above -- mentioned accretion -- induced col- lapse scenario because all the existing calculations assume that the composition of the He -- exhausted core is carbon -- free. Nevertheless, these authors studied only the mass -- conservative evolutionary tracks for the relevant range of stellar masses, whereas in a close binary system the com- position of the final remnant could be dramatically af- fected by the previous evolutionary phase if the star is interacting with a companion. In any case, progress in the right direction has been made, but further exploration is still worthwhile. In this paper we follow the evolution of a 10 M⊙ model star of solar metallicity belonging to a close binary sys- tem, from its main sequence phase until an oxygen -- neon core develops. In order to keep consistency with our pre- vious results, for the calculations reported in this paper we have used the same evolutionary code described in Ri- tossa, Garc´ıa -- Berro & Iben (1996). In particular we have not considered overshooting, as was done there. This pro- cedure might lead to somewhat smaller cores than in the case in which overshooting is considered, but we expect the differences to be small. Different mass loss episodes are caused by the presence of a close companion and are followed in detail. Specifically, the model loses most of its hydrogen -- rich envelope in a case B Roche -- lobe overflow episode. It is worth mentioning at this point that we have taken special care in treating the mass zoning and the time steps during the mass loss episodes and the carbon burning phase (about 2 000 lagrangian mesh points are used during the most problematic phases). Moreover, the distribution of mesh points is regularly updated at each time step. Our algorithm puts mesh points where they are most needed (that is, where the gradients of the physical variables are strong) and eliminates them where they are not necessary. Once this is done, the values of the physical variables are interpolated at the new mesh points in order to properly compute the temporal derivatives. If necessary the mesh can be updated at each iteration. As is usually done, throughout this work we will refer to the star that is initially more massive as the primary component whereas the secondary component will be its companion. We have assumed that the initial orbital pa- rameters and the mass of the secondary component al- low the whole evolution to proceed without reversal mass transfer and without disruption or merger events. This poses some constraints on the mass of the secondary which will be discussed later. The plan of the paper is the following. In the second section of this paper we explain our choice of the initial orbital parameters and the assumed scenario. In the third section we present a description of the overall evolution of our model star until carbon ignition sets in, and we com- pare it with the evolution of an isolated star of the same mass. This section is also devoted to the study of the re- sulting mass loss episodes. In the fourth section we briefly describe the evolution during the carbon burning phase and we discuss the final characteristics of the remnant. Finally our major findings and conclusions are described in section §5. 2. The scenario In this section we set up our evolutionary scenario and we describe our choices for the initial orbital parameters. The reader should keep in mind that our main goal is to pro- vide a successful scenario to test the formation of massive ONe white dwarfs in close binary systems that could ulti- mately lead either to a cataclysmic variable or to collapse to neutron star dimensions through the accretion -- induced collapse alternative. Therefore our choice of the initial pa- rameters of the binary system is effectively influenced by the desired final outcome. A possible observational coun- terpart of the proposed scenario could be the binary sys- tem IK Peg (HR 8210, HD 204188), which has an orbital period of 21.7 days, and it is composed of a massive white dwarf of mass MWD ≃ 1.15 M⊙ and a main sequence star of 1.4 M⊙ (Smalley et al., 1996). 4 Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs Fig. 1. Approximate evolution of the orbital parameters of the binary system and outline of the possible final outcomes. The dashed line divides the calculations reported in this paper and the three possible outcomes. Our starting point is a primordial system composed of a 10 M⊙ star and its <∼ 5 M⊙ companion, which we will refer to as the primary and secondary components respectively (see Fig. 1). The system undergoes a case B mass transfer, this being the most likely possibility, and so the primary starts losing mass after the onset of hydrogen burning in a shell. Unlike early case -- B mass loss episodes, late case -- B mass transfer has been little studied up to now. The reasons for this are multiple but perhaps the most important one is the additional computational diffi- culties that arise when studying such a phase. However, it has been recently pointed out (Tauris & Dewi, 2001) that the actual definition of the resulting core of the primary after the mass loss episode could influence the final orbital parameters. We consider it interesting, therefore, to study a late case -- B mass loss episode. This determines the range of values for the initial Roche lobe radius of the primary, that we actually choose by considering the evolutionary track for the single 10 M⊙ model star followed by Garc´ıa -- Berro et al. (1994), and we keep it constant at 210 R⊙ dur- ing the whole process (see the discussion below). Nonethe- less, it is worth mentioning that we have conducted a series of numerical experiments in which the Roche lobe radius has been changed to values as small as 120 R⊙ and we have not found significant differences. Given the initial mass relation and the Roche lobe of the primary, we can obtain the orbital separation between both components by appliying the equation (Eggleton, 1983): = 0.46q2/3 , 0.6q2/3 + ln (1 + q1/3) RL A where RL denotes the effective radius of the Roche lobe of the primary, A indicates the orbital separation and q is the mass ratio (M1/M2) between the components. The initial orbital separation turns out to be A ∼ 150−350 R⊙ (1) Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs 5 and the orbital period is Porb ∼ 0.1 − 1.0 yr, depending on the initial mass of the companion (between 1 and 5 M⊙), which are reasonable values. As it will be shown in detail in the next section, a number of uncertainties are involved in the first mass loss episode. These uncertainties are basically caused by the formation of a common envelope. This common envelope is formed when the primary star attains the highest values for its mass loss rates, and most probably removes a large amount of mass and angular momentum from the system. Consequently, the evolution of the orbital parameters can- not be accurately followed, and this can only be done in an approximate way. In order to determine the mass lost by the primary, we assume that all the matter overflowing the Roche lobe is lost by the star, and we simply stop the process when its radius dramatically decreases below that of the Roche lobe (see §3.1). The second mass loss episode starts when the surface radius of the primary again exceeds that of its Roche lobe. In this case, this is due to the onset of helium burning in the shell. This process is most likely conservative both in mass and in angular momentum and, again, stops when the radius of the primary decreases below the value of the Roche lobe. After this temporary decrease in radius, car- bon ignition continues and the primary starts the ascent along the Super -- AGB but does not fill again its Roche lobe. Once carbon is exhausted in the innermost regions, it is a stellar wind rather than Roche lobe overflow that induces mass loss and, thus, we use the parametrization of Nieuwenheuzen and de Jager (1990) in order to account for the mass loss rates. As happened during the first mass loss episode, there are several uncertainties involved in this process, which can have consequences on the final fate of the system. In fact, depending on the choice of the effi- ciency of the winds, the resulting orbital parameters can vary significantly (Umeda et al. 1999). If the system sur- vives the Super -- AGB phase, we are left with an ONe white dwarf, as a result of the evolution of the primary, and a main sequence star that, at some time, will fill its Roche lobe and give rise to reversal mass transfer. Even though we do not mean to get deep into the study of the probability of occurrence of the different final out- comes, we will briefly outline the different possibilities for the final stages of the life of the binary in terms of the mass transfer rates during the reversal mass transfer. De- pending on this mass transfer, the final outcome can be either a cataclysmic variable if it is lower than a critical rate, a supernova explosion if it is larger or an ONe white dwarf if it is much larger. A more detailed study can be found in §5. 3. Evolution until the beginning of the carbon burning phase and main mass loss episodes The presence of a close companion has several conse- quences for the evolution of our 10 M⊙ model star. In particular, there are two Roche -- lobe overflow mass loss episodes. The first one occurs just after the main sequence phase when our model star reaches red giant dimensions, and the second one happens shortly after the exhaustion of central helium. Figure 2 shows the evolution of our model in the Hertzsprung -- Russell diagram. Times to evolve to each la- beled point along the evolutionary track are given in the second column of Table 1, where we also provide the most important characteristics of these particular models for the hydrogen and helium burning phases. Also shown in the last columns of table 1 are the mass of the primary and secondary stars (M1 and M2), and the expected or- bital period (Porb). The solid lines represent the evolu- tionary phases during which no mass loss occurs, whereas the dotted lines correspond to the evolutionary phases where mass loss occurs: from B to C, for the first mass loss episode, and from shortly after K to L, for most of the second Roche -- lobe overflow. Note, however, that the final part of the second Roche -- lobe overflow occurs when carbon has already been ignited in the core (see §3.4 and §4) and, thus, it is not shown in figure 2. As expected, the effects of the mass loss episodes considerably modify the evolutionary track when compared with the evolution of an isolated star of the same initial mass (Garc´ıa -- Berro & Iben 1994). For instance, although the descent along the red giant branch takes place at a slightly higher tempera- ture than in the case of an isolated star (4 400 K instead of 4 300 K), it is followed by a sudden shift to bluer re- gions of the diagram due to the adiabatic expansion and cooling that accompanies the first mass loss episode (see below). The second major different feature is that the de- velopement of the blue loop during the core helium burn- ing phase takes place at lower luminosities than in the case of the isolated model star (6.9 × 103 L⊙ instead of 7.9 × 103 L⊙). In Figure 3 we show the temporal evolution of the ra- dius of the star (top panel) and of the convective regions associated with nuclear burning (bottom panel), from the zero age main sequence up to the off -- center ignition of carbon in the helium -- exhausted core. The inserts show with higher resolution the evolution of the convective re- gions and of the radius during the two mass -- loss episodes. The initial central convective zone of Fig. 3 is due to the high fluxes engendered by the CN -- cycle reactions and per- sists until hydrogen vanishes at the center. An off -- center convective region forms later, very much in the same way as in the case of a single star. This convective region is due to the release of gravitational potential energy dur- ing the overall contraction phase that follows the exhaus- tion of central hydrogen and the establishement of the hydrogen -- burning shell that occurs from points C to D in Fig. 2. The thick solid line in the lower panel corre- sponds to the total mass of the primary which, as it can be seen, decreases dramatically at the begining of the first dredge -- up -- process, which occurs simultaneously with the 6 Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs Table 1. Characteristics of models at various points in the H -- R diagram of Fig. 2. Model t (1014s) 6.1785 6.1795 6.1893 6.2917 6.2936 6.2954 6.3555 6.8505 7.2076 7.2122 7.2178 7.2638 log L log Teff 3.79 4.12 4.16 3.99 3.96 3.98 3.84 3.68 3.90 3.82 3.89 4.21 3.67 3.63 3.64 3.64 3.59 3.59 3.62 4.18 3.72 3.81 3.64 3.73 log R log ρc 6.67 2.07 2.32 6.62 6.51 2.32 6.45 2.23 6.45 2.32 6.44 2.32 2.41 6.43 6.47 2.01 7.42 2.02 7.43 1.85 2.18 6.43 7.66 2.18 log Tc MHe 1.97 8.16 8.16 1.97 1.97 8.16 2.04 8.18 2.04 8.18 2.04 8.18 8.19 2.05 2.05 8.24 2.05 8.15 2.05 8.42 8.45 2.06 2.06 8.60 A B C D E F G H I J K L MC 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.76 0.86 0.91 0.93 1.04 M1 10.0 M2 1.0 − 5.0 Porb (yr) 0.1 − 1.0 2.4 1.7 − 7.0 0.3 − 0.5 1.9 2.2 − 9.3 0.1 − 1.0 Surface abundances after the first dredge -- up. SS stands for single star. XO XN XC Table 2. Model 10 M⊙ (SS) 10 M⊙ (CBS) 9 M⊙ (SS) XH 0.681 0.664 0.696 X3 1.23 × 10−5 1.13 × 10−5 1.49 × 10−5 XHe 0.305 0.324 0.291 1.79 × 10−3 1.64 × 10−3 1.85 × 10−3 2.83 × 10−3 3.46 × 10−4 2.48 × 10−3 7.71 × 10−3 7.19 × 10−3 8.03 × 10−3 Fig. 2. Evolutionary track of the primary component in the Hertzsprung -- Russell diagram. The physical quantities of the labeled models are shown in Table 1. first Roche -- lobe overflow. Since this mass loss episode oc- curs in the presence of a deep convective envelope, the associated time scales are short. On the contrary, we will see that the second mass -- loss episode is much more stable since it is not associated with a dredge -- up episode. This behaviour makes the two mass -- loss episodes completely different and has important consequences. For instance, we expect to find a different pattern of surface composi- tion after the first mass -- loss episode and dredge -- up, when compared to the evolution of a single star. The surface composition of both model stars and that of the isolated 9 M⊙ star just after the end of the dredge -- up episode can Fig. 3. Evolution of the main structure parameters as a function of time during the first and the second mass -- loss episodes. The upper panel shows the evolution of the radius of the primary. The inserts show the mass -- loss episodes with higher resolution. The lower panel shows the convective zones engendered during the evolution up to off -- center ignition of carbon. be found in Table 2. As we shall show below, the compar- ison with the 9 M⊙ star is relevant for this study since some of the results obtained in the calculations reported here are much closer to the isolated 9 M⊙ model than to those of the 10 M⊙ single star. As can be seen, the helium and nitrogen contents are significantly higher in the case of a model star belonging to a close binary system. Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs 7 3.1. The first mass loss episode Our model undergoes a Case B mass loss process that starts shortly after hydrogen combustion in a shell has been established, and the surface radius reaches the Roche Lobe radius which, as previously mentioned, we have adopted to be 210 R⊙. In computing this mass -- loss episode we have (somewhat arbitrarily) assumed that the primary keeps a constant radius which is coincidential with the Roche -- lobe radius and that all the overflowing matter will be lost by the primary. Although this is a classical prescription there are, of course, other alterna- tives (Nelson & Eggleton 2000). It is nevertheless worth noticing that when using the last approximation with sec- ondary masses of around ∼ 5 M⊙, the Roche lobe radius changes by ∼ 10% and, consequently, we do not expect this to have a large impact on our results. We thus defer such study to a forthcoming publication. Fig. 4. Global characteristics of the primary component as a function of time during the first mass -- loss episode. The upper panel shows the total mass as a function of time. The lower panel shows the surface radius and the luminosities provided by hydrogen and helium burning. The vertical thin lines delimit the two phases of mass loss. The surface luminosity is mostly provided by hydrogen combustion. In the first insert in the top and bottom panels of Fig. 3 the first mass -- loss episode is shown. Also, in Figure 4 we show the evolution of several interesting quantities like the hydrogen and helium nuclear luminosities. This mass -- loss episode occurs in two distinct phases. The first and most violent phase occurs as the convective envelope is still advancing to the interior of the star and when helium has already been ignited at the center, leading to the for- mation of a central convective region, which reaches its equilibrium value when log(LHe/L⊙) ≃ 3.3. This phase is clearly marked by the two arrows in the top insert of Fig. 3 and by the thin vertical lines of Fig. 4. During this first phase, the bulk of the hydrogen -- rich envelope is lost (∼ 7.6 M⊙), leaving a remnant of 2.4 M⊙ of which ∼ 0.5 M⊙ corresponds to the remaining H -- rich convective envelope. The duration of this phase is 3.0 × 104 yr. Since the dependence of the variation of the total ra- dius of the star on the mass is ∆R ∼ RM −1/3 (de Loore & Doom 1992), the mass -- loss process experiences a posi- tive feedback. Therefore, during the very first part of this mass -- loss episode the feedback of the process allows very high values for the mass loss rates, that can reach val- M1 ∼ 10−3 M⊙yr−1. At these very high ues as high as mass -- loss rates the star is no longer able to keep a con- stant radius and at the end of this phase the radius of the primary falls below the value of the Roche -- lobe, and thus the mass transfer onto the secondary temporarily stops. As the evolution continues, the primary again fills its Roche -- lobe radius, leading to a subsequent phase of mass -- loss. During this second phase only ∼ 0.05 M⊙ are lost by the primary. Thus, a small portion of the H -- rich envelope remains even at the end of the first mass -- loss episode. The time scale for this second phase is signifi- cantly longer (1.6 × 105 yr), leading to much more modest mass -- loss rates: M1 ∼ 10−5 M⊙yr−1. The existence of a deep convective envelope surround- ing the H -- exhausted core of the primary and, conse- quently, the high values of the resulting mass -- loss rates very much enhance the possibility that the system is em- bedded in a common envelope. No definite and accurate study of this kind of structure has been performed up to now. Thus, the part of the process in which a common en- velope forms is plagued with many uncertainties, the most important one perhaps being the influence of the common envelope on the orbital parameters. The expressions pro- vided by Eggleton (1983) and Paczy´nski (1971) can only be taken as approximations, or even upper limits, and it is necessary to rely on estimates that have been obtained for other systems that presumably have undergone a com- mon envelope phase, like the cataclysmic variable U Gem. These estimates yield values for the angular momentum losses that can be as high as 70 -- 90% (Iben & Tutukov 1984). In any case, when a common envelope is present, the angular momentum losses are expected to be high due to the release as sound waves of the energy generated by frictional drag in the matter of the envelope. There- fore, from an initial separation between the components of Ai ∼ 350−450 R⊙ (correspondig to 1 M⊙ ≤ M2 ≤ 5 M⊙ respectively) one can assume a typical 80% angular mo- mentum loss and apply the equation J 2 orb = GAi (M1M2)2 M1 + M2 (2) to get the final separation between components after the first Roche -- lobe overflow, Af ∼ 150−350 R⊙. It should be noted, nevertheless, that this way of estimating the final 8 Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs orbital parameters usually leads to larger orbital separa- tions than the values obtained when using the treatment of Webbink (1984). We can estimate the ratio between the mass accreted by the secondary and the mass lost by the primary, β, by M2 = assuming stable accretion onto the secondary. Thus, M2/τKH, where τKH is the Kelvin -- Helmholtz timescale. Considering typical values for the luminosity and for the surface radius of a main sequence companion of mass be- tween 1 and 5 M⊙, we get a reasonable value for the mass M2 ∼ 10−4M⊙. Taking accretion rate of the secondary, into account that most of the mass is lost by the primary during the first 3.0 × 104 yr, we can easily calculate an average value for β during this mass loss episode, which turns out to be β ∼ 0.1. Of course, it can be argued as well that another reasonable way to estimate β is to adopt the Eddington limit instead of the thermal timescale, this being a firm upper limit. For the range of relevant param- eters, this procedure would lead in our case to accretion rates 3 times larger, or equivalently, to β ∼ 0.3. Neverthe- less it should be noted that the calculations of Hjellming & Taam (1991) predict a significantly lower value of β, of the order of 0.01. Thus, the mass accreted by the secondary will be in the range 0.7 − 2.0 M⊙, at most. Since the main sequence lifetime of a 7.0 M⊙ star is t ∼ 9.0 × 1014 s it is clear that reversal mass transfer will not be enabled until carbon is exhausted in the inner core of the primary (see Table 1 and §4 below), in accordance with our scenario. Fig. 5. Opacity profiles in the hydrogen -- rich envelope during the first Roche lobe overflow. The models displayed correspond to the different times labeled in Fig. 4. The in- sert shows, with higher resolution, the opacity profiles for the last models. We have considered the possibility that part of the matter ejected by the primary component might be ac- creted back by the same star. In order to obtain some hints of the possible outcomes to this problem, we have estimated the duration of the common envelope phase as in Iben & Tutukov (1984) and compared it with the ther- mal time scale of the primary. As the latter is much longer than the expected time for the common envelope to re- main bound to the system, there are strong reasons to admit that there is not enough time for the ejected gas to cool down, lose kinetic energy and be overtaken by the gravitational potential of the primary. It is also worth noting that the duration of the first dredge -- up episode is longer in the calculations reported here than in the case of the isolated model described in Garc´ıa -- Berro & Iben (1994). This is due to the fact that the fast release of gravitational energy during the com- pression phase that happens in the middle of the first mass loss episode cannot be evacuated solely by radia- tion and, thus, allows convection to persist. On the other hand, one should not forget that the overall evolutionary time scales are longer for decreasing masses and this ef- fect would also have an influence on the duration of the dredge -- up episode. In order to find an explanation that, at least partially, accounts for the behaviour of the envelope during the first mass loss episode, in Fig. 5 we show the opacity profiles of several specific models for the times labeled in Fig. 4 (models a to g). The fastest expansion phase is coinci- dential with models a to c, in which the existence of a deep convective envelope, basically composed of hydrogen at relatively low temperature, leads to a high opacity and, thus, to an inefficient transport of the energy. Therefore, an important portion of the energy generated at the hy- drogen burning shell is not driven outwards but, instead, it is transformed into internal energy in the envelope and, then, into work of expansion, thus keeping the value of the surface radius very close to the Roche lobe radius. When the model reaches approximately 4 M⊙ (model d), a large portion of the hydrogen -- rich envelope is al- ready lost, and the inner and hotter layers are exposed. At this point, on the one hand we have less mass able to absorb the flow of energy and, on the other, the opacity is also smaller. Consequently, both phenomena allow nuclear energy to flow almost freely to the surface without being transformed into work of expansion and, at model e, when the mass of the star is about 2.45 M⊙, a fast overall (al- most adiabatic) contraction of the convective envelope oc- curs, leading the surface radius to drop below the value of the Roche lobe and, hence, mass loss temporarily halts. Fi- nally, for model f the opacity increases due to compression and, thus, a new phase of expansion occurs which drives the surface radius to again reach the value of the Roche lobe radius. Thus the mass loss episode is restored for a brief interval, until some more cool hydrogen -- rich layers Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs 9 are lost. Finally the opacity definitely decreases (model g) and the mass loss episode is finished. It is worth mentioning at this point that we get a remnant of slightly higher mass than that of the models found in the existing literature (Iben & Tutukov 1985; Dom´ınguez, Tornamb´e & Isern 1994). This is due to the fact that in our model the hydrogen -- rich envelope is not completely lost during the first Roche lobe over- flow episode. However, since the rest of the remaining hydrogen -- rich envelope is lost later during the second Roche lobe overflow episode, and since the growth of the helium core is very small during the time between the two mass loss episodes (approximately 0.08 M⊙), we ex- pect that the influence on the CO and ONe core sizes and compositions is negligible and so the final results will not be substantially different. 3.2. The second mass loss episode outer envelope, from which the matter is removed, energy is transported by radiation instead of being transported by convection, allows the process to take place in a more stable way than in the first mass loss episode. Hence, the mass loss rates are much smaller ( M1 ∼ 10−6 M⊙yr−1) during most of the process, except at two phases, during which the values of the mass loss rate are one order of magnitude higher. The first phase corresponds to the be- ginning of the mass loss episode, when the remainder of the hydrogen -- rich envelope is lost, and the second phase happens nearly at the end of the process, when carbon is ignited off -- center. The end of the process is determined by a steep decrease in the surface radius of the star that takes values below the Roche lobe radius. The moderate values we get for the mass loss of the pri- mary support the hypothesis of conservative mass trans- fer. Furthermore, one can compare the luminosity associ- ated with accretion onto the secondary, which is given by Lacc = M2(ΦL1 − ΦR2 ), where ΦL1 and ΦR2 stand for the gravitational potential computed at the Roche lobe ra- dius and at the radius of the secondary respectively, with the Eddington luminosity of the secondary, LEdd. For a set of typical values for the secondary star, we get that Lacc ≪ LEdd and, thus, according to Han et al. (1999) conservative mass transfer is most likely. 3.3. Overall characteristics before carbon ignition Fig. 6. Global characteristics of the primary component as a function of time during the second mass loss episode. The upper panel shows the total mass as a function of time. The lower panel shows the surface radius and the lu- minosities provided by hydrogen, helium and carbon burn- ing. The vertical thin lines delimit the second mass loss episode. The surface luminosity is mostly provided by he- lium combustion. The second mass loss episode starts when helium burn- ing begins in a shell and, as a consequence, its surface radius approaches again that of the Roche lobe (see Fig. 3), which now is ∼ 80 R⊙. The global characteristics (the mass and radius and the luminosities associated with the active burning regions) of the primary during the second mass loss episode are shown in Fig. 6. The fact that in the Fig. 7. Evolution in the log ρ − log T plane of the central layer of the isolated 10 M⊙ star (dashed line) and the of the primary component of the close binary system studied here (solid line), until the onset of carbon burning. In Fig. 7 we show the evolution of the center of the pri- mary in the log ρ − log T diagram from the main sequence 10 Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs until the onset of carbon ignition (solid line), and the evo- lution of the center of the isolated 10 M⊙ star (dashed line). As one can expect from the previous discussion, the differences between both cases start to show up at the on- set of helium burning at the center, which, as explained above, is almost coincident with the first mass loss episode. The evolution of the primary component of the close bi- nary system leads to higher central densities for the same value of the temperature than its isolated counterpart. In fact, mass loss has two obvious consequences. Firstly, the size of the central helium -- burning convective region is smaller for the case studied here than in the evolution of isolated 10 M⊙ star. Secondly, the mass of the primary component of the binary system is much smaller than its isolated counterpart. Accordingly, the He -- exhausted core of the primary is smaller (∼ 1.05 M⊙) and with higher central densities than the core resulting from the evolu- tion of a single 10 M⊙ star (∼ 1.15 M⊙) and very similar to that of a 9 M⊙ isolated star (∼ 1.04 M⊙). The abundances in the central regions of the core are also affected by mass loss in a similar way, and so, the resulting composition is more similar to that of the single 9 M⊙ model. At the end of the second mass loss episode the mass of the primary is 1.88 M⊙, of which the inner- most 1.05 M⊙ corresponds to a carbon -- oxygen core which is surrounded by a helium -- rich envelope. The chemical composition profiles at this moment are shown in Fig. 8. As can be seen in this figure the abundance profiles of our model before carbon ignition reveal a higher carbon and neon content, and lower amounts of oxygen and mag- nesium than those of the single 10 M⊙ model. Also the helium burning shell is narrower in the case studied here, due to the fact that the (helium) envelope over this burn- ing shell is much smaller. 4. The carbon burning phase The carbon burning phase takes place under conditions of partial degeneracy and -- as one can expect from the com- parison between the characteristics of the primary star in a close binary system and that of single star models of similar mass before carbon burning -- many similarities are also found when analysing this phase of the evolution. Actually, the most prominent features, such as the carbon flashes and the convective regions associated with each one of these (see Fig. 9) are very similar for the case stud- ied here and those of a 9 M⊙ model star. These flashes reach about 107 − 108 L⊙, and last for approximately 103 yr. However, as it happens in the case of the evolution of isolated stars of this mass range, during most of the car- bon burning phase the surface luminosity remains almost constant and close to a value of log(L/L⊙) = 4.3. The first flash is a prototypic one. As in the case of iso- lated stars within this mass range, carbon is ignited off -- center due to neutrino cooling of the central regions. The mass coordinate at which carbon is ignited is ∼ 0.42 M⊙, Fig. 8. Abundance profile of the core of the primary at the end of the second mass loss episode, which is practically coincidential with the beginning of carbon burning. Fig. 9. Evolution of the nuclear luminosities during the carbon -- burning phase (upper panel), and of the associated convective regions (lower panel). The solid and the dotted lines in the upper panel correspond, respectively, to the carbon and helium luminosities. very close to the value obtained for the single 9 M⊙ star (∼ 0.43 M⊙). As carbon luminosity increases, most of the energy generated by nuclear reactions is used to increase the temperature of the adjacent layers, thus forcing high temperature gradients, which ultimately lead to the for- mation of a convective zone that allows a more efficient transport of energy. This increase in the temperature is Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs 11 Fig. 10. Relevant structural and dynamical quantities for a model at time t = 7.277763 × 1014 s, just before the dis- appearance of the convective envelope, in a region extend- ing from below the helium -- burning shell until the surface of the star. Panel a: velocity (solid line) and helium profile (dotted line). Panel b: time variation of the temperature (solid line) and density (dotted line). Panel c: energy re- lease rates (nuclear and gravothermal) along with the time derivative of the internal energy and the work of expan- sion. The nuclear energy release rate has been divided by 102 in order to fit into the scale. The thin solid line marks the position of inner edge of the convective envelope. followed by an expansion of the nearby layers and in par- ticular, of the helium burning shell, which cools down and, hence, its luminosity temporarily decreases. Apart from the formation of the inner convective shell, the flash also affects the location of the inner edge of the convective en- velope that moves deeper into the star when the carbon luminosity increases and receeds when the carbon lumi- nosity decreases again, very much in the same fashion as in the isolated 9 M⊙ star. The second flash occurs at a smaller mass coordi- nate (carbon is re -- ignited at the point where the pene- tration of the inner edge of the previous convective shell was maximum) and the physical mechanisms operating are very similar to those of the first one. However, its maximum strength is considerably smaller (∼ 107 L⊙ in- stead of ∼ 108 L⊙). After the most violent phase of this flash is over, the carbon luminosity does not decrease be- low that of helium but, rather, keeps a stationary value, log(LC/L⊙) ≃ 4.6, which is slightly larger than the value of the helium luminosity, log(LHe/L⊙) ≃ 4.3. During Fig. 11. Same as figure 10, but for time t = 7.278391 × 1014 s, just after the convective episode. this phase the carbon -- burning flame propagates inwards at a roughly constant speed -- see Garc´ıa -- Berro et al. (1997) for a detailed description of the energy balance established during this phase -- reaching the center at t ≃ 7.282 × 1014 s. A very distinctive feature of the case studied here when compared to the evolution of an isolated 9 M⊙ star is that at t ≃ 7.278 × 1014 s the outer convective envelope disap- pears as the carbon burning front advances to the center. In fact, the second dredge -- up in the isolated 9 M⊙ star is caused by the expansion and cooling of the layers just below the base of the convective envelope (Garc´ıa -- Berro et al. 1997), which allows the increase of the radiative temperature gradient in this region and, so, the inner ad- vance of convection down to Mr ≃ 1.2 M⊙. After that, the base of the convective envelope remains at an approx- imately constant position, as the energy supplied by the helium burning shell and the opacity of the nearby lay- ers can keep the temperature gradient high enough. In the case studied here, the base of the convective envelope reaches the same position as in the single star due to a similar mechanism, but, unlike the case of the isolated 9 M⊙ star, this position cannot be maintained, instead, it receeds and disappears. The reason for this behaviour can be explained with the help of figures 10 and 11, where we show some relevant structural and dynamical quantities for times t = 7.277763 × 1014 s, just before the disappear- ance of the convective envelope, and t = 7.278391×1014 s, just after the convective episode. 12 Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs Panel c of figure 10 shows that there are three inter- esting and distinct regions in the star. The first of these regions is below the helium discontinuity and there, the work of expansion fed by carbon burning is devoted ba- sically to lift degeneracy and, thus, we have expansion at almost constant temperature (panel b). On top of this re- gion we have the helium burning shell where the energy supplied by nuclear reactions is used to build up the lumi- nosity profile (panel a). In the region between the helium burning shell and the base of the convective envelope, the flux is partially transformed into heating and, at the same time, the whole region is collapsing. This in turn causes a large temperature gradient at Mr ≃ 1.7 M⊙. As the evo- lution continues, the temperature in this region steadily increases. Thus, the temperature gradient ultimately flat- tens and the inner edge of the convective envelope con- sequently receeds. Panel c of figure 11 shows that the nuclear energy released by the helium burning shell re- mains the same, but now most of this energy merely flows outwards (panel a), rather than being transformed into work of expansion. Instead, the whole region on top of the helium burning shell is compressed leading to a non -- homogeneous increase of the temperature (panel b) which effectively erases the temperature gradient and forces the disappearance of the convective region. This translates into an increase in the luminosity for Mr >∼ 1.4 M⊙. When carbon -- burning in the central regions is over, a series of small shell burning episodes occurs (Fig. 9). Each one of these episodes has a decreasing strength and leads to the almost complete exhaustion of carbon in a core of 1.05 M⊙. It is interesting to note that, unlike what hap- pens with the rest of these mild flashes, the first of these is not accompanied by a substantial decrease in the he- lium luminosity. The reason for this behaviour is twofold. Firstly, the duration of this flash is smaller and, hence, less energy is released. Secondly, and most important, there exists a thick radiative layer (of about 0.7 M⊙) between the initial location of the convective carbon -- burning shell and the helium -- burning shell. Therefore, the energy is al- most totally absorbed before reaching the helium -- burning shell and, thus, no expansion nor cooling of the helium burning shell are produced. During all this phase the core contracts, and the outer layers of the star begin a new expansion as the outer convective envelope reappears and its inner edge advances to the interior of the star. How- ever, the radius of the star remains below the Roche lobe radius and only exceeds its value at the end of the second of these flashes. Since, as discussed above, the second mass loss episode was most probably conservative and during the bulk of the carbon burning phase in the central regions (say t <∼ 7.282 × 1014 s) the model star keeps its radius be- low that of the Roche lobe, the orbital parameters remain unaffected. As explained before, only at the very end of the carbon burning phase the flash activity is accompa- nied by a rapid increase of the surface radius of the star Fig. 12. Final abundances of the primary star after the carbon burning phase. Note the existence of a carbon -- rich buffer on top of the carbon exhausted core beyond the Roche lobe radius. This, of course, translates into a new mass loss episode for t >∼ 7.283 × 1014 s. This mass loss episode lasts for a short period of time, of about ∼ 2×1011 s and the mass loss rates are of ∼ 10−4 M⊙ yr−1 at the beginning of the episode, becoming much smaller as the evolution proceeds. After this short phase of Roche lobe overflow, the radius of the primary star turns out to be very large, but its rate of increase slows down con- siderably. At this evolutionary stage, the density of the extended envelope is so small that a stellar wind could also drive the loss of the remaining envelope. In any case, stellar winds play a significant role which can be compara- ble to Roche lobe overflow. Thus, we assume that a stellar wind removes mass from the surface following closely the parametrization of Nieuwenheuzen and de Jager (1990) which gives typical rates of ∼ 5 × 10−7 M⊙ yr−1 (this is the rate considered in Fig. 9). However, since these mass loss rates are poorly known, we have conducted a series of numerical experiments where we have changed the mass loss rate from 10−5 M⊙ yr−1 to 10−7 M⊙ yr−1 and we have found essentially the same results for the very final part of the carbon burning phase, except of course for the mass of the remaining helium envelope. The exhaustion of central carbon is not complete but, instead, at the innermost 0.3 M⊙ there remains remains a small amount of unburnt carbon, reaching a maximum abundance of XC = 0.025 at Mr = 0.25 M⊙ (see Fig. 12). Analogous to what was found in the series of pa- pers devoted to the evolution of isolated stars within this mass range, most of the ashes of carbon burning are sodium, neon and oxygen, the abundance of magnesium being very small. This could have important consequences Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs 13 for the accretion -- induced collapse scenario (Guti´errez et al., 1996). It is also noteworthy that the relatively thick carbon -- rich buffer is located just below the helium enve- lope, which could make the resulting He -- rich white dwarf practically indistinguishable from an observational point of view from a regular massive carbon -- oxygen white dwarf (Weidemann 2000). 5. Summary and discussion We have followed the evolution of a 10 M⊙ primary com- ponent of solar metallicity belonging to a close binary sys- tem with a secondary of low or intermediate -- low mass, from its main sequence phase until carbon is exhausted in the core and a degenerate remnant is formed. Our main goal consisted of extending the previously existing calcu- lations of this kind of system in order to follow in full detail the carbon burning phase, which at present has not yet been well studied. We have found that the primary star undergoes three mass -- loss episodes. The first of these episodes occurs after hydrogen exhaustion in the core and most problably is not conservative, due to the existence of a deep convective envelope which leads to the formation of a common envelope. The second happens after helium exhaustion, and it is most problably conservative. Finally, the third one occurs when carbon burning in a series of succesive shells sets in. We have analysed the changes in the structure and composition that the primary star suf- fers while simultaneously undergoing carbon burning and mass loss. The determination of the mass loss rates at this very late stage of the evolution, whether due to Roche lobe overflow, or to stellar winds, is still an open question, and therefore, we have explored a broad range of mass loss rates ranging from 10−7 to 10−5M⊙ yr−1 in order to un- derstand how the uncertainties affect the gross properties of the resulting degenerate object. We have found that our incomplete knowledge of the mass -- loss rate does not affect the final result as far as the core is concerned. In particu- lar, carbon is not reignited. Thus, the final characteristics of the remnant are well determined. After the evolution through these different burning stages and under the in- fluence of a close companion, the initially 10 M⊙ star will form a ∼ 1.1 M⊙ degenerate remnant, with an oxygen -- neon core of ∼ 0.9 M⊙ surrounded by a carbon -- oxygen rich mantle of ∼ 0.2 M⊙ and a thin helium envelope. We have found remarkable differences between our re- sults and those found by other authors that have also followed the evolution of a 10 M⊙ model star up to the carbon burning phase. For instance, Nomoto (1982, 1984) obtained a final core of larger mass (∼ 1.3 M⊙) with signif- icantly higher values for the abundances Ne20 and Mg24, and lower values for the O16 and Na23 abundances. The ultimate reason for this difference is that the nuclear reac- tion rates used by Nomoto (Fowler, Caughlan & Zimmer- mann, 1975) are different from ours (Caughlan & Fowler, 1988). In this regard, it is important to recall here that in this paper we have chosen the same physical inputs as in Ritossa, Garc´ıa -- Berro & Iben (1996) in order to remain consistent with our previous calculations. Since then there have been some new determinations of the thermonuclear reaction rates, particularly the NACRE compilation (An- gulo et al., 1999). However, for the most important reac- tion channels involved in carbon burning, the differences are not expected to be large and, thus, the chemical com- position of the core can be considered as relatively safe, perhaps the most important expected difference being an even smaller amount of Mg24 (Palacios et al., 2000; Jos´e, Coc & Hernanz, 1999). The comparison with the ONe core obtained by Dom´ınguez et al. (1993) it is not easy either, since they start from a different initial CO core prior to carbon burning. Moreover, they removed the outer enve- lope during the carbon burning phase and, hence, their treatment for this phase differs substantially from ours. This, in particular, could be the reason for the central re- gion of unburnt carbon which is quite apparent in their results, although poor mass resolution of the central re- gions -- see the discussion in Garc´ıa -- Berro et al. (1997) -- cannot be totally discarded. Regarding the average abun- dances of the ONe core, they obtained, using similar nu- clear reaction rates, X(O16) = 0.72, X(Ne20) = 0.25 and X(Mg24) = 0.03. The high value of the O16 abundance is surprising, whereas the rest of the abundances are simi- lar. Moreover, they did not find Na23, which is an impor- tant isotope in our calculations. Thus, all these differences could be due to the use of a simplified nuclear network, much smaller than ours. In summary, the composition of our ONe core differs from the former results, especially for the Mg24, Na23 and Ne20 nuclei, which are the elements onto which electron captures happen and, consequently, may substantially affect the determination of the explo- sive ONe ignition density. The fact that the mass of the ONe core found by these authors (∼ 0.93 M⊙) is quite similar to that obtained here is, however, encouraging. The definite final result is not necessarily an oxygen -- neon white dwarf, because the evolution of the entire sys- tem has not stopped yet. In fact, as the secondary com- ponent evolves, it could reach such a dimension that its radius could exceed that of its Roche lobe, and mass trans- fer could take place onto the remnant. Depending on the evolution of the orbital parameters and, ultimately, on the mass transfer rates, different possibilities arise: 1. A cataclysmic variable could be formed, if the ac- cretion rates are below the critical rates for hy- drogen burning. At typical accretion rates of about 10−9M⊙ yr−1 onto such a white dwarf, the amount of hydrogen -- rich material from the secondary that must be accreted before an outburst occurs ranges between 10−4 -- see, for instance, Jos´e & Hernanz (1998) and references therein -- and 10−5M⊙ (Schwartzman et al., 1994). If we accept the results of Iben and Fujimoto (1992) that significant hydrogen abundance extends in 14 Pilar Gil -- Pons & Enrique Garc´ıa -- Berro: Formation of ONe white dwarfs a mixture with the degenerate material down to a dis- tance of about 10−5M⊙ under the surface, and that approximately 2 × 10−5M⊙ are expelled during each outburst, we conclude that between 3 000 and 4 000 outburts (in a total time of at least 3 -- 4 Myr) must occur before the white dwarf is definitely deprived of its 0.035 M⊙ CO rich layer, and so, before significant amounts of Ne20 can be detected in the expelled ma- terial. 2. If stable accretion is possible, the degenerate object can increase its mass up to the Chandrasekhar mass, fi- nally leading to a weak supernova explosion, activating the accretion -- induced collapse scenario and leaving a neutron star. A possible observational counterpart for this outcome could be the massive radio pulsar PSR J0045-7319 (Kaspi et al. 1994), which belongs to a bi- nary system in which its companion is an intermediate mass (M >∼ 4M⊙) B star. More evolved systems may include an intermediate mass binary system consist- ing of a pulsar and a white dwarf. In this sense, up to seven pulsars with these characteristics have recently been discovered (Edwards & Bailes, 2001). In particu- lar, we consider PSR J1756-5322 as a very likely coun- terpart because the mass of the white dwarf component is >∼ 0.55M⊙. 3. Finally, if the mass transfer rates are so large that the white dwarf is not able to accrete all the mass lost of its companion, a common envelope would form again, and most of the mass would be lost by the system, leaving an ONe white dwarf belonging to a close bi- nary system. A possible observational counterpart to this outcome is IK Peg (Smalley et al., 1996), which is a confirmed binary system composed of an unevolved star plus a massive white dwarf (MWD ≃ 1.15 M⊙), possibly made of oxygen and neon. Acknowledgements. Part of this work was supported by the Spanish DGES project number PB98 -- 1183 -- C03 -- 02, by the MCYT grant AYA2000 -- 1785, by the CIRIT and by Sun Mi- croSystems under the Academic Equipment Grant AEG -- 7824 -- 990325 -- SP. We also wish to thank J. Jos´e for carefully reading the manuscript and to our referee (P.P. Eggleton) for his very valuable comments and suggestions. References Angulo, C. Arnould, M., Rayet, M., Descouvemont, P., Baye, D., Leclercq -- Willain, , C., Coc, A., Barhoumi, S., Aguer, P., Rolfs, C., Kunz, R., Hammer, J.W., Mayer, A., Pa- radellis, T., Kossionides, S., Chronidou, C., Spyrou, K., Degl'Innocenti, S., Fiorentini, G., Ricci, B., Zavatarelli, S., Providencia, C., Wolters, H., Soares, J., Grama, C., Rahighi, J., Shotter, A., & Lamehi Rachti, M., 1999, Nucl. Phys. A, 656, 3 Caughlan, G.R., & Fowler, W.A., 1988, Atom. Data & Nuc. Data Tables, 40, 283 De Loore, C., & de Greve, J.P., 1992, A&A Suppl. Ser., 94, 453 De Loore, C., & Doom, C.W.H., 1992, "Structure and evolution of single and binary stars", Astrophysics and Space Science Library, vol. 179 (Dordrecht: Kluwer) De Loore, C., & Vanbeveren, D., 1992, A&A, 260, 273 De Loore, C., & Vanbeveren, D., 1994, A&A, 292, 463 De Loore, C., & Vanbeveren, D., 1995, A&A, 304, 220 Dom´ınguez, I., Tornamb´e, A., & Isern, J., 1993, ApJ, 419, 268 Edwards, R.T., & Bailes, M., 2001, ApJ, 547, L37 Eggleton, P.P., 1983, ApJ, 268, 368 Fowler, W.A., Caughlan, G.R., & Zimmermann, B.A., 1975, ARA&A, 13, 69 Fujimoto, M.Y., & Iben, I., 1992, ApJ, 399, 646 Garc´ıa -- Berro, E., & Iben, I., 1994, ApJ, 434, 306 Garc´ıa -- Berro, E., Ritossa, C., & Iben, I., 1997, ApJ, 485, 765 Guti´errez, J., Garc´ıa -- Berro, E., Iben, I., Isern, J., Labay, J., & Canal, R., 1996, ApJ, 459, 701 Han, Z., & Webbink, R. F., 1999, A&A, 349, L17 Han, Z., Tout, C.A., & Eggleton, P.P., 2000, MNRAS, 319, 215 Hjellming, M.S., Taam, R.E., 1993, ApJ, 370, 709 Iben, I., & Tutukov, A., 1984, ApJ Suppl., 54, 335 Iben, I., & Tutukov, A., 1985, ApJ Suppl., 58, 661 Iben, I., Ritossa, C., & Garc´ıa -- Berro, E., 1997, ApJ, 489, 772 Jos´e, J., Coc, A., & Hernanz, M., 1999, ApJ, 520, 347 Jos´e, J., & Hernanz, M., 1998, ApJ, 494, 680 Kaspi, V.M., Johnston, S., Bell, J.F., Manchester, R.N., Bailes, M., Bessell, M., Lyne, A.G., & D'Amico, N., 1994, ApJ, 423, L43 Langer, N., Deutschmann, A., Wellstein, A., & Hoflich, P., 2000, A&A, 362, 1041 Miyaji, S., Nomoto, K., Yokoi, K., & Sugimoto, D., 1980, PASJ, 32, 303 Nelson, C.A., & Eggleton, P.P., 2001, ApJ, in press (astro- ph/0010269) Nieuwenhuijzen, H., & de Jager, C., 1990, A&A, 231, 134 Nomoto, K., 1984, ApJ, 277, 791 Palacios, A., Leroy, F., Charbonnel, C., & Forestini, M., 2000, in "The Galactic Halo: from globular clusters to field stars", ed.: A. Noels, P. Magain, D. Caro, E. Jehin, G. Parmentier, & A. Thoul, p. 67 Paczy´nski, B., 1971, ARA&A, 9, 183 Ritossa, C., Garc´ıa -- Berro, E., & Iben, I., 1995, ApJ, 460, 489 Ritossa, C., Garc´ıa -- Berro, E., & Iben, I., 1999, ApJ, 515, 381 Saio, H., & Nomoto, K., 1998, ApJ, 500, 388 Smalley, B., Smith, K.C., Wonnacott, D. & Allen, C.S., 1996, MNRAS, 278, 688 Swartzman, E., Kovetz, A., & Prialnik, D., 1994, MNRAS, 269, 323 Tauris, T.M., & Dewi, J.D.M., 2001, A&A, in press Umeda, H., Nomoto, K., Yamaoka, H., & Wanajo, S., 1999, ApJ, 513, 861 Van der Linden, T.J., 1987, A&A, 178, 170 Webbink, S., 1984, ApJ, 277, 355 Weidemann, V., 2000, A&A, 363, 647 Woosley, S.E., Weaver, T.A., & Taam, R.E., 1980, in "Texas Workshop on Type I Supernovae", ed. J.C. Wheeler (Austin: Univ. of Texas Press), 96 Whyte, C.A., & Eggleton, P.P., 1980, MNRAS, 190, 801 Whyte, C.A., & Eggleton, P.P., 1985, MNRAS, 214, 357
astro-ph/0310904
2
0310
2004-02-10T16:04:24
Generalized phantom energy
[ "astro-ph", "gr-qc", "hep-ph", "hep-th" ]
We examine cosmological models with generalized phantom energy (GPE). Generalized phantom energy satisfies the supernegative equation of state, but its evolution with the scale factor is generally independent, i.e. not determined by its equation of state. The requirement of general covariance makes the gravitational constant time-dependent. It is found that a large class of distinct GPE models with different evolution of generalized phantom energy density and gravitational constant, but the same equation of state of GPE have the same evolution of the scale factor of the universe in the distant future. The time dependence of the equation of state parameter determines whether the universe will end in a de Sitter-like phase or diverge in finite time with the accompanying "Big Rip" effect on the bound structures.
astro-ph
astro-ph
IRB-TH-6/03 4 0 0 2 b e F 0 1 2 v 4 0 9 0 1 3 0 / h p - o r t s a : v i X r a Generalized phantom energy H. Stefanci´c∗ Theoretical Physics Division, Rudjer Boskovi´c Institute, P.O.Box 180, HR-10002 Zagreb, Croatia Abstract We examine cosmological models with generalized phantom energy (GPE). Generalized phantom energy satisfies the supernegative equation of state, but its evolution with the scale factor is generally independent, i.e. not determined by its equation of state. The requirement of general covariance makes the gravitational constant time-dependent. It is found that a large class of distinct GPE models with different evolution of generalized phantom energy density and gravitational constant, but the same equation of state of GPE have the same evolution of the scale factor of the universe in the distant future. The time dependence of the equa- tion of state parameter determines whether the universe will end in a de Sitter-like phase or diverge in finite time with the accompanying "Big Rip" effect on the bound structures. Results of recent cosmological observations, such as distant supernovae of type Ia (SNIa) [1] and cosmic microwave background radiation (CMBR) [2], have dra- matically altered our perception of the dynamics and composition of the universe and reshaped the landscape of standard cosmology [3]. The universe seems to be in the phase of accelerated expansion, which started at a relatively small redshift, z ∼ 1. This acceleration is attributed to a new form of matter, usually referred to as dark energy, the nature of which is still not definitely established. Observations ∗[email protected] 1 indicate that the energy density of the universe is very close to its critical density where dark energy presently accounts for approximately 2/3 of the total energy density, while the remaining 1/3 comes predominantly from dark matter, another unidentified component of the universe. The most prominent and studied candi- dates for the title of dark energy are the cosmological constant [4, 5, 6] (together with its dynamical variants, such as renormalization group running cosmological constant [7, 8, 9]), quintessence [10] and the Chaplygin gas [11]. The majority of dark energy models share a common constraint on their equa- tion of state (pd and ρd represent pressure and energy density of dark energy, respectively) pd = wρd , (1) where w ≥ −1. Such a constraint is, however, not justified by the unbiased fits to the data of cosmological observations. Moreover, the allowed interval for the pa- rameter of the equation of state extends significantly into the region with w < −1. The use of observational data on CMBR, large scale structure (LSS), SNIa and Hubble parameter measurements from the Hubble Space Telescope (HST) un- der the assumption of the redshift independent parameter w give the restriction −1.38 < w < −0.82 at the 95% confidence level [12]. Therefore, a possible su- pernegative equation of state of dark energy deserves due attention. A new type of dark energy with the equation of state characterized by w < −1 was proposed in [13] and named phantom energy. Phantom energy is considered to be separate from other components of the universe and its energy-momentum tensor is conserved separately. In such a setting, the equation of state of dark energy determines its evolution with the scale factor a. The supernegative nature of the equation of state of the phantom energy leads to the growing energy density of phantom energy ρd ∼ a−3(1+w), for a constant parameter w. The cosmological dynamics of the universe with such a phantom energy component possesses many interesting features [14]. The growth of the energy density of phantom energy drives the scale factor of the universe to infinity in finite time. The increasing negative pressure of phantom energy leads to the unbounding of all bound structures in the universe. This dramatic and picturesque scenario of the cosmic doomsday was ap- propriately named "Big Rip". The formulation of microscopic models for phantom energy [15] relies on the machinery developed in quintessence models, namely the evolution of the scalar field in a suitably chosen potential. However, the description of phantom energy may require an introduction of some nonstandard alterations, e.g. the negative kinetic term of the scalar field. Detailed considerations of the Lagrangians describing phantom energy show that in some cases the universe with phantom energy ends in a "Big Rip", while in others it asymptotically approaches the de Sitter expansion. In this paper, we consider models with generalized phantom energy (GPE). First, we set up a more general model of the evolution of the universe with phantom energy. We assume that there are two components of the universe: the dark energy component (which will have the phantom energy characteristics), and the "ordinary" matter component with the respective energy densities ρd and ρm. The "ordinary" matter is taken to satisfy the equation of state pm = γρm , 2 (2) where γ ≥ 0. Furthermore, we assume that the energy-momentum tensor of the "ordinary" matter is conserved T µν m;ν = 0 . (3) The equation given above ensures that the parameter of the equation of state governs the evolution of the "ordinary" matter energy density, i.e. ρm = ρm,0(cid:18) a a0(cid:19)−3(1+γ) . Dark energy has the equation of state pd = wρd , (4) (5) where w generally depends on time explicitly or implicitly, via explicit dependence on some other time-dependent quantity, such as the scale factor a. In the case of dark energy, we allow the possibility of non-conservation of the energy-momentum tensor, i.e. T µν d;ν 6= 0 . (6) Thus, the evolution of the dark energy density is not determined by the parameter from its equation of state. With the properties of the components of the universe defined, we can specify the laws of its evolution. We start from the Einstein equation Gµν = −8πGT µν , (7) d + T µν where Gµν is the Einstein tensor and T µν = T µν m is the total energy- momentum tensor. The reconciliation of the requirement of the general covari- ance of (7) and the non-conservation relation (6) is possible with the promotion of gravitational constant G into a space-time dependent quantity. This change can be interpreted as a modification of the dynamics of General Relativity. This additional dynamics is effectively described by the introduction of space-time dependence of G. We consider the models where G is a function of time only, G = G(t). Mod- els with the time-dependent G were extensively studied in the framework of the time-dependent cosmological term Λ(t) [16]. The covariant derivative of (7) then implies This equation can be rewritten in the form (G(t)T µν );ν = 0 . d(G(ρm + ρd)a3) = −G(pm + pd)da3 . (8) (9) Combining the evolution laws (4) and (9) and introducing w ≡ −1 + κ (where κ describes the deviation from the parameter of the equation of state inherent to the cosmological constant) we arrive at G(ρm + ρd) + G ρd + 3κHGρd = 0 . (10) Here H = a/a is the Hubble parameter, while dots denote time derivatives. Equa- tion (10) clearly shows the generality of the model. In the case of the constant G, 3 we recover the standard equation of conservation of T µν d . Equation (10) shows that the time evolution of G is the result of two competing effects. Namely, for dark energy with growing energy density, the second term in (10) causes the decrease of G, while for negative κ, the third term in (10) increases G with time. Finally, Friedmann equations for the evolution of the scale factor complete the set of evolution equations (4) and (10) a(cid:19)2 (cid:18) a + k a2 = 8π 3 G(ρm + ρd) , a a = − 4π 3 G(ρm + ρd + 3pm + 3pd) . (11) (12) The set of equations (4), (10) and (11) reveals that we have essentially two independent equations for three dynamical quantities G, ρd and a (assuming that κ is the function of these quantities and time). Without a more specific identification of the dynamics of G or ρd, it is not possible to solve the aforementioned set of equations. However, as we show below, with mild assumptions about the evolution of dark energy with the scale factor, it is possible to obtain information on the future evolution of the universe for general G and ρd satisfying the equations given above. Next, we introduce the concept of generalized phantom energy (GPE). Gener- alized phantom energy is the form of dark energy satisfying the equation of state (1) with the non-conserved energy-momentum tensor (6) and the following two properties: (a) GPE energy density is a non-decreasing function of the scale factor, (b) GPE equation of state satisfies κ ≤ 0. We further examine the future evolution of the universe. In the sufficiently distant future we have ρm ≪ ρd and ρm can be neglected in the evolution equations. Equations (10) and (11) thus become and a(cid:19)2 (cid:18) a + k a2 = 8π 3 Gρd d dt (Gρd) + 3κHGρd = 0 . Furthermore, from equation (14), we obtain d(Gρd) Gρd = −3κ da a . As the condition −κ ≥ 0 is satisfied by assumption (b), we obtain Gρd ≥ (Gρd)0 . (13) (14) (15) (16) Therefore, as Gρd is a growing function in an expanding universe, for large a we can disregard the term k/a2 in equation (13). For the flat universe, this approximation is exact, while for the closed or the open universe, this approximation is applicable in the sufficienly distant future. 4 Finally, we end up with the following two equations for the dynamics of the universe in the distant future: H 2 = 8π 3 (Gρd) , d dt (Gρd) + 3κH(Gρd) = 0 . (17) (18) By combining equations (17) and (18), we obtain an equation for the evolution of the Hubble parameter H with time with the solution H(t) = dH dt + 3 2 κH 2 = 0 , 1 + 3 H(t0) 2 H(t0)R t t0 κ(t′)dt′ (19) (20) . Once we have found the expression for the evolution of the Hubble parameter, it is easy to obtain an expression for the evolution of the scale factor a a(t) = a(t0)exp Z t t0 dt′ H(t0) t0 κ(t′′)dt′′! . 1 + 3 2 H(t0)R t′ (21) General solutions (20) and (21) exhibit some interesting features. The evolution of the universe in the sufficiently distant future is governed only by the parameter of the equation of state of dark energy. The precise form of the growth of ρd with the scale factor a is irrelevant in this limit. This implies that the entire class of models with different functional forms of ρd and G, obeying the same equation of state, show the same behaviour in the sufficiently distant future. Therefore, we can divide all GPE models with the characteristics specified above into classes with the same equation of state. An important question regarding the fate of the universe is whether, for a particular class of generalized phantom energy models, a and H diverge in finite time or reach infinite values only in infinite time. For the Hubble parameter H, the answer is straightforward. There will be no divergence of H in finite time if the denominator of the expression on the right-hand side of (20) remains positive for all times. This leads to the condition Z ∞ t0 (−κ(t′))dt′ < 2 3H(t0) . (22) R ∞ As in this case there is no singularity in H(t) in finite time, the scale factor a(t) also does not diverge in finite time. In order to have the convergence of the integral t0 (−κ(t′))dt′ required in (22), the function κ(t) has to tend to zero at asymp- totically large times. Therefore, for generalized phantom matter which exhibits no divergence of H or a in finite time, the parameter of the equation of state approaches −1, i.e. generalized phantom energy approaches the time-dependent cosmological term. In the case when the condition (22) is not satisfied, the Hubble parameter H diverges in finite time t. From Friedmann equations we have a = Ha , 5 (23) a = (cid:18)1 − 3 2 κ(cid:19) H 2a . (24) These expressions indicate that, when H diverges in finite time t, both a and a diverge as well, so the scale factor a cannot remain finite, but diverges in finite time t as well. From the general expessions (20) and (21), we can obtain evolution laws for the conceptually simple, but important case [13] κ(t) = −κ0 . (25) With such a choice for the parameter of the equation of state of generalized phan- tom energy, we have the following evolution laws: H(t) = H(t0) 1 − 3 2 H(t0)κ0(t − t0) , a(t) = a(t0)(cid:18)1 − 3 2 H(t0)κ0(t − t0)(cid:19)− 2 3κ0 . (26) (27) These solutions clearly show the onset of the divergence in H and a. The universe with generalized phantom energy with the constant parameter of the equation of state evolves to infinity in finite time. Comparison with the case of the "standard" phantom energy [13, 14] shows that, for the same parameter of the equation of state κ(t), the scale factor follows the same evolution law. Given the fact that the parameter of the equation of state does not determine the scaling with a, and that G is variable in the framework of generalized phantom energy, it is by no means obvious that coincidence of this sort should exist. However, from the equation (10), we readily see that for the case of constant G, we recover the equation of evolution for the "standard" phantom energy. As far as the evolution in the sufficiently distant future is concerned, the "standard" phantom energy model is just one instance of the class of generalized phantom energy models with the same function κ(t). Given the same evolution properties of the broad class of GPE models with the same κ(t), it is natural to look at the destiny of bound structures, another peculiarity of phantom energy models [14]. The relevant quantity with respect to the stability of the bound structures is the analogue of the gravitational potential proportional to the quantity G(ρd + 3pd) = (−2 + 3κ)Gρd. Equation (17) shows that Gρd ∼ H 2 and Gρd grows with time. If the condition (22) is not satisfied, H and Gρd diverge in finite time. Furthermore, as ρd grows with the scale factor, Gρd certainly increases compared to G. For gravitationally bound systems, the GPE contribution of the order ∼ G(ρd + 3pd)R3 (where R denotes the characteristic spatial scale of the bound system) overwhelms the "mass" contribution ∼ GM (M denotes the mass of the bound system). Gravitationally bound systems fall apart in finite time. For the systems bound by electromagnetic or strong forces, mere growth of Gρd ensures their unbounding at some finite time before the time at which scale factor goes to infinity. Consequently, all bound structures are unbound in finite times. The scenario of the "Big Rip" is present in generalized phantom energy models as well. 6 Finally, let us make some comments on fundamental aspects of the GPE model. As the gravitational constant G(t) is time-dependent, the description of the grav- itational sector in the GPE model represents a declination from the Einsteinian gravity. One important aspect is whether the scale factor a really describes the growth of length scales. One can raise two arguments in favour of the standard interpretation of the scale factor a. The first is that no intervention in the geo- metrical structure or interpretation of the left-hand side of equation (7) has been made. The other, more physical one, is that the density of "nonrelativistic" matter scales as ρm ∼ a−3 in our GPE model, equation (4) with γ = 0. Given that no interaction (production or annihilation) of the "ordinary" matter component with other components is assumed, this fact establishes a as a natural measure of the growth of length scales. In some theories with the time-dependent effective gravitational constant, such as scalar-tensor or nonminimally coupled scalar field theories, one can construct many mathematically equivalent theories using conformal transformations. It turns out that all these theories are not physically equivalent, i.e. some formulations are more physically viable than others (the Einstein frame formulation is more viable than the Jordan frame formulation) [17]. Generally, it might be of interest to consider conformally related models of GPE obtained by the transformation of the type gµν = f (G(t))gµν , where f is a suitably chosen function. However, the time variation of G(t) in our model can be very general and includes possibilities to which requirements on the choice of the conformal frame do not necessarily apply. Some examples of such a variation are the renormalization group running of G [7, 8, 9] or the time variation of G emanating from extra dimensions [18]. In conclusion, in this letter we have considered cosmological models with the time-dependent gravitational constant G and dark energy with the supernegative equation of state (phantom energy). Phantom energy is generalized in the sense that its equation of state does not determine its evolution with the scale factor a, i.e. GPE density becomes an independent function of the scale factor. The require- ment of general covariance in this setting imposes conditions on the gravitational constant G which acquires time dependence. Investigation of future dynamics of the generalized phantom energy models with growing generalized phantom energy density and the parameter of the equation of state less than −1 exhibits some general properties. A large class of models with different evolutions of ρd and G, but the same equation of state of GPE, have the common law of the evolution of the scale factor a in the sufficiently distant future. The time dependence of the GPE parameter of the equation of state determines whether the universe evolves infinitely in a de Sitter regime or diverges in finite time. One would expect that bounds on the variation of G in the past epochs of the evolution of the universe would produce the most stringent constraints on the parameters of the GPE model. Therefore, it is important to point out that our main results qualitatively do not depend on the size of the parameter κ or on the intensity of growth of ρd (of course, within classes of these parameters that satisfy or do not satisfy the con- dition (22)). For smaller parameter values and slowlier varying functions ρd and G, the onset of the general evolution (dependent only on κ) will come later. For instance, for constant and negative κ, but very small κ, the entire class of GPE models leads to the "Big Rip" event, but at very late times. 7 Clearly, the present accelerating phase of the evolution of the universe carries the seed of the possibly very dramatic future of our cosmos. Therefore, more precise observations of the past variation of ρd and G with time (redshift) will be able to unravel the fate of the universe. Acknowledgements. The author would like to thank N. Bili´c, B. Guberina and R. Horvat for useful comments on the manuscript. This work was supported by the Ministry of Science and Technology of the Republic of Croatia under the contract No. 0098002. References [1] A.G. Riess et al., Astron. J. 116 (1998) 1009; S. Perlmutter et al., Astrophys. J. 517 (1999) 565. [2] P. de Bernardis et al., Nature 404 (2000) 955; A.D. Miller et al. Astrophys. J. Lett. 524 (1999) L1; S. Hanany et al., Astrophys. J. Lett. 545 (2000) L5; N.W. Halverson et al., Astrophys. J. 568 (2002) 38; B.S. Mason et al., astro-ph/0205384; D.N. Spergel et al., astro-ph/0302209; L. Page et al. astro-ph/0302220. [3] W.L. Freedman, M.S. Turner, astro-ph/0308418; S.M. Carroll, astro-ph/0310342. [4] S. Weinberg, Rev. Mod. Phys. 61 (1989) 1. [5] P.J.E. Peebles, B. Ratra, Rev. Mod. Phys. 75 (2003) 559. [6] T. Padmanabhan, Phys. Rept. 380 (2003) 235. [7] A. Babic, B. Guberina, R. Horvat, H. Stefancic, Phys. Rev. D 65 (2002) 085002; B. Guberina, R. Horvat, H. Stefancic, Phys. Rev. D 67 (2003) 083001. [8] I.L. Shapiro, J. Sola, Phys. Lett. B 475 (2000) 236; I.L. Shapiro, J. Sola, JHEP 0202 (2002) 006; I.L. Shapiro, J. Sola, C. Espana-Bonet, P. Ruiz-Lapuente, Phys. Lett. B 574 (2003) 149. [9] A. Bonanno, M. Reuter, Phys. Lett. B 527 (2002) 9; E. Bentivegna, A. Bo- nanno, M. Reuter, astro-ph/0303150. [10] B. Ratra, P.J.E. Peebles, Phys. Rev. D 37 (1988) 3406; P.J.E. Peebles, B. Ratra, Astrophys. J. 325 (1988) L17; C. Wetterich, Nucl. Phys. B 302 (1988) 668; R.R. Caldwell, R. Dave, P.J. Steinhardt, Phys. Rev. Lett. 80 (1998) 1582; I. Zlatev, L. Wang, P.J. Steinhardt, Phys Rev. Lett. 82 (1999) 896. [11] A. Yu. Kamenshchik, U. Moschella, V. Pasquier, Phys. Lett. B511 (2001) 265; N. Bilic, G.B. Tupper, R.D. Viollier, Phys. Lett. B 535 (2002) 17; M.C. Bento, O. Bertolami, A.A. Sen, Phys. Rev. D 66 (2002) 043507. [12] A. Melchiorri, L. Mersini, C.J. Odman, M. Trodden, Phys. Rev. D 68 (2003) 043509. [13] R.R. Caldwell, Phys. Lett. B 545 (2002) 23. [14] R.R. Caldwell, M. Kamionkowski, N.N. Weinberg, Phys. Rev. Lett. 91 (2003) 071301. 8 [15] C. Armendariz-Picon, T. Damour, V. Mukhanov, Phys. Lett. B 458 (1999) 209; T. Chiba, T. Okabe, M. Yamaguchi, Phys. Rev. D 62 (2000) 023511; V. Faraoni, Int. J. Mod. Phys. D 11 (2002) 471; S.M. Carroll, M. Hoffman, M. Trodden, Phys. Rev. D 68 (2003) 023509; S. Nojiri, S.D. Odintsov, Phys. Lett. B 562 (2003) 147; S. Nojiri, S.D. Odintsov, Phys. Lett B 565 (2003) 1; P. Singh, M. Sami, N. Dadhich, Phys. Rev. D 68 (2003) 023522; G.W. Gibbons, hep-th/0302199; L.P.Chimento, R. Lazkoz, gr-qc/0307111; J.G. Hao, X.Z. Li, Phys. Rev. D 68 (2003) 083514; M.P.Dabrowski, T. Stachowiak, M. Szyd- lowski, hep-th/0307128; J.G. Hao, X.Z. Li Phys. Rev. D 67 (2003) 107303; J.G. Hao, X.Z. Li, astro-ph/0309746; V. Faraoni, gr-qc/0307086. [16] A. Beesham, Nuovo Cimento 96 B (1986) 17; A. Beesham, Int. J. Theor. Phys. 25 (1986) 1295; A-M.M. Abdel-Rahman, Gen. Rel. Grav. 22 (1990) 655; M.S. Berman, Phys. Rev. D 43 (1991) 1075; M.S. Berman, Gen. Rel. Grav. 23 (1991) 465; R.F. Sistero, Gen. Rel. Grav. 32 (1991) 1265; D. Kalligas, P. Wesson, C.W.F. Everitt, Gen. Rel. Grav. 24 (1992) 351; T. Singh, A. Beesham, Gen. Rel. Grav. 32 (2000) 607; A.I. Arbab, A. Beesham, Gen. Rel. Grav. 32 (2000) 615; A.I. Arbab, Spacetime and Substance 1(6) (2001) 39; A.I. Arbab, astro-ph/0308068; J. Ponce de Leon, gr-qc/0305041. [17] V. Faraoni, E. Gunzig, P. Nardone, Fund. Cosmic Phys. 20 (1999) 121; V. Faraoni, E. Gunzig, Int. J. Theor. Phys. 38 (1999) 217. [18] P. Loren-Aguilar, E. Garcia-Berro, J. Isern, Yu.A. Kubyshin, Class. Quant. Grav. 20 (2003) 3885. 9
astro-ph/9808189
1
9808
1998-08-19T02:49:20
The Multi-Phase Medium in the Interstellar Complex N44
[ "astro-ph" ]
We have obtained high-resolution HI observations of N44, one of the largest HII complexes in the Large Magellanic Cloud. The distribution and internal motions of the HI gas show dynamic effects of fast stellar winds and supernova blasts. Numerous HI holes are detected, with the most prominent two corresponding to the optically identified superbubbles Shell 1 and Shell 2. The HI gas associated with Shell 1 shows an expansion pattern similar to that of the ionized gas shell, but the mass and kinetic energy of the HI shell are 3--7 times those of the ionized gas shell. The total kinetic energy of the neutral and ionized gas of Shell 1 is still more than a factor of 5 lower than expected in a pressure-driven superbubble. It is possible that the central OB association was formed in a molecular cloud and a visible superbubble was not fully developed until the ambient molecular gas had been dissociated and cleared away. This hypothesis is supported by the existence of a molecular cloud toward N44 and the fact that the apparent dynamic age of the superbubble Shell 1 is much shorter than the age of its OB association LH47. Accelerated HI gas is detected at the supernova remnant 0523-679. The mass and kinetic energy in the associated HI gas are also much higher than those in the ionized gas of 0523-679. Studies of interstellar gas dynamics using ionized gas alone are clearly inadequate; neutral gas components must be included.
astro-ph
astro-ph
Appear to ApJ Vol. 503 (Aug 20) The Multi-Phase Medium in the Interstellar Complex N44 Sungeun Kim,1 You-Hua Chu,2 Lister Staveley-Smith,3 and R. Chris Smith4 ABSTRACT We have obtained high-resolution H i observations of N44, one of the largest H ii complexes in the Large Magellanic Cloud. The distribution and internal motions of the H i gas show dynamic effects of fast stellar winds and supernova blasts. Numerous H i holes are detected, with the most prominent two corresponding to the optically identified superbubbles Shell 1 and Shell 2. The H i gas associated with Shell 1 shows an expansion pattern similar to that of the ionized gas shell, but the mass and kinetic energy of the H i shell are 3 -- 7 times those of the ionized gas shell. The total kinetic energy of the neutral and ionized gas of Shell 1 is still more than a factor of 5 lower than expected in a pressure-driven superbubble. It is possible that the central OB association was formed in a molecular cloud and a visible superbubble was not fully developed until the ambient molecular gas had been dissociated and cleared away. This hypothesis is supported by the existence of a molecular cloud toward N44 and the fact that the apparent dynamic age of the superbubble Shell 1 is much shorter than the age of its OB association LH47. Accelerated H i gas is detected at the supernova remnant 0523−679. The mass and kinetic energy in the associated H i gas are also much higher than those in the ionized gas of 0523−679. Studies of interstellar gas dynamics using ionized gas alone are clearly inadequate; neutral gas components must be included. Subject headings: ISM: bubbles - ISM: supernova remnant - Magellanic Clouds - X-ray, HI, HII: ISM 1Mount Stromlo and Siding Spring Observatories, ANU, Weston Creek PO, Canberra, ACT 2611, Australia 2Astronomy Department, University of Illinois at Urbana-Champaign, Urbana, IL 61801 3Australia Telescope National Facility, CSIRO, P.O. Box 76, Epping, NSW 2121, Australia 4Astronomy Department, University of Michigan, Ann Arbor, MI 48109-1090 -- 2 -- 1. Introduction N44, cataloged by Henize (1956), is a luminous H ii complex in the Large Magellanic Cloud (LMC). It contains an assortment of compact H ii regions, filaments, and shells of all sizes, as well as three OB associations, LH47, 48, and 49 (Lucke & Hodge 1970). As shown in Figure 1, N44 is dominated by a prominent shell around LH47 in the central region (Chu & Mac Low 1990). In the surroundings are: diffuse H ii regions and filaments to the east; compact H ii regions encompassing LH49 to the south; a large faint shell to the west; faint filaments to the north; a luminous, compact H ii region around LH48 on the northeastern rim of the main shell; and numerous small, bright, single-star H ii regions in the outskirts of N44. Previous studies of ionized gas in N44 have yielded many interesting results. The main shell and the large shell to the west are designated as Shell 1 and Shell 2, respectively, and their expansion patterns have been studied by Meaburn & Laspias (1991). Shell 1 has been modeled as a superbubble using energy input implied by the observed massive stellar content; it is shown that the expansion velocity of Shell 1 is much higher than expected (Oey & Massey 1995). Diffuse X-ray emission has been detected in N44, indicating the existence of 106 K gas (Chu & Mac Low 1990; Wang & Helfand 1991). The X-ray emission within Shell 1 and Shell 2 has been suggested to be generated by supernova remnants (SNRs) shocking the shell walls; the outward extension of X-ray emission at the southern periphery of Shell 1 has been interpreted as a breakout; and the bright diffuse X-ray source at 6′ − 7′ northeast of Shell 1 has been identified as a new SNR (Chu et al. 1993; Magnier et al. 1996). Following the naming convention of using truncated B1950 coordinates, this remnant will be named SNR 0523−679. Interactions between massive stars and the interstellar medium (ISM) have led to the derivation of a multi-phase ISM model (McKee & Ostriker 1977) and have been used to explain the disk-halo connection in a galaxy (Norman & Ikeuchi 1989). The physical structure of N44 vividly demonstrates the dynamical interaction between massive stars and the ISM. Thus, N44 provides an excellent laboratory to study in detail the distribution, physical conditions, and relationship of the different phases of the ISM in a star forming region. Previous studies of N44 have concentrated on only the warm and hot ionized gas. In order to carry out a comprehensive investigation of the multi-phase structure of N44, we have recently obtained high-resolution H i observations of N44 with the Australia Telescope Compact Array (ATCA) and high-dispersion, long-slit echelle spectra of N44 with the 4m telescope at Cerro Tololo Inter-American Observatory. In this paper we report these observations (§2), describe the interaction between H i and H ii gases (§3), analyze the H i gas associated with Shell 1 and Shell 2 (§4), and study the acceleration of H i gas by a SNR (§5). A discussion is given at the end (§6). -- 3 -- 2. Observations 2.1. H i Observations The H i data of N44 are extracted from the H i aperture synthesis survey of the LMC made with the Australia Telescope Compact Array (ATCA), which consists of 6 22-m antennas. The details of this survey are given by Kim et al. (1997) and preliminary results are presented by Kim and Staveley-Smith (1997). In summary, the observations were made in four configurations: 750A on 1995 February 23 to March 11, 750B on 1996 January 27 to February 8, 750C on 1995 October 15 to 31, and 750D on 1994 October 26 to November 9. These configurations each have five antennas with a maximum baseline of 750 m. The combined configuration has 40 independent baselines ranging from 30 to 750 m, with a baseline increment of 15.3 m. The resultant angular resolution is 1.′0 for the H i images presented in this paper. The largest angular structure that the images are sensitive to is ∼ 0.◦6. The observing band was centered at 1.419 GHz with a bandwidth of 4 MHz. The band was divided into 1024 channels; these channels were re-binned into 400 channels after an on-line application of Hanning smoothing and edge rejection. The final data cube used here has a velocity coverage of 190 to 387 km s−1, and a velocity resolution of 1.65 km s−1. All H i and Hα velocities reported in this paper are heliocentric. The data cube of N44 was extracted from the full mosaic of the LMC, which consisted of 1344 pointing centers. Approximately four pointing centers were included in the data cube of N44. The phase and amplitude calibrators were PKS B0407−658 for some fields and PKS B0454−810 for the others. The primary ATCA calibrator, PKS B1934−638 (assumed flux density 14.9 Jy at 1.419 GHz), was observed at the start and end of each observing day. These observations served both for bandpass calibration and flux density calibration. The data were edited, calibrated, and mosaicked in miriad. The pixel size is 20′′ and the size of the N44 cube is 23 × 23 × 120. Superuniform weighting (Sramek and Schwab 1989) was applied to the uv data with an additional Gaussian taper. The resulting data were then Fourier transformed and mosaicked in the image plane. The maximum entropy method was used to deconvolve this cube. The final cube was constructed by convolving the maximum entropy model with a Gaussian of FWHM 1′ × 1′. The final rms noise, measured in line-free parts of the final cube, is 9K. 2.2. Echelle Observations To examine the kinematic properties of the ionized gas in N44, we obtained high- dispersion spectroscopic observations in the Hα+[N ii] lines with the echelle spectrograph -- 4 -- on the 4m telescope at Cerro Tololo Inter-American Observatory on 1995 January 18. Using a post-slit Hα filter and replacing the cross-disperser with a flat mirror, we were able to observe both Hα and [N ii]λλ6548, 6583 lines over a long slit. The 79 lines mm−1 echelle grating and the red long focus camera were used. The data were recorded with a Tek 2048×2048 CCD. The 24 µm pixel size corresponds to about 0.82 A (3.75 km s−1 at the Hα line) along the spectral direction and 0.′′267 along the spatial direction. The spectral coverage was limited by the filter width to 125 A. The spatial coverage was limited by the slit length to ∼ 4′, with some vignetting in the outer 1.′5. The slit width was 250 µm, or 1.′′66, and the instrumental FWHM was 16.1±0.8 km s−1. The wavelength calibration and geometric distortion correction were carried out with the use of Th-Ar lamp exposures in the beginning of the night. The observations were sufficiently deep that the geocoronal Hα component was detected in each frame. The geocoronal Hα component (at zero observed velocity) provided an accurate and convenient reference for velocity measurements. The observed velocities were then converted to heliocentric velocities. The echellograms are shown in Figure 1. Only Hα line is presented in the echellograms. The [N ii] lines are detected at a much lower S/N ratio than the Hα line, hence they are not presented here. 2.3. Optical Imaging In order to compare the neutral gas components to the ionized gas in the N44 region, we obtained optical Hα emission-line images with a CCD camera mounted at the Newtonian focus of the Curtis Schmidt telescope at CTIO on 18 February 1994 UT. The detector was a front-illuminated Thomson 1024 × 1028 CCD with 19µ pixels, giving a scale of 1.′′835 pixel−1 and a field of view of 31.′3. A narrow-bandpass Hα filter (λc = 6561 A, ∆λ = 26 A) was used to isolate the emission, and a red continuum filter (λc = 6840 A, ∆λ = 95 A) was used to obtain images of the continuum background (images in other emission lines were also obtained, but those will be discussed in another paper). Multiple frames were obtained through each filter, amounting to total integration times of 1200s in Hα and 600s in the continuum. The data were reduced with IRAF5, and multiple frames were shifted and combined to obtain the images shown in Figures 1 and 3b. 5IRAF is distributed by the National Optical Astronomy Observatories (NOAO). -- 5 -- 3. Interaction between H i and H ii Components The integrated H i map of N44 (Figure 2) shows two clear depressions that correspond to the interiors of the optically defined Shell 1 and Shell 2, and peaks that are adjacent to Hα features. The relationship between the H i gas and the H ii gas is clearly complex. The ATCA channel maps of H i (Figure 3a) kinematically resolve the distribution of H i along the line of sight, and allow more precise identification of physical structures in the H i distribution. Comparing these iso-velocity maps of H i to the Hα image (Figure 3b), we may examine the relationship between the H i and H ii gas in greater detail and accuracy. From the comparisons of H i and Hα images, we have found indications of: (1) H i holes and expanding shells, (2) acceleration of H i, (3) compression of H i, and (4) the opening of a breakout in H i. These phenomena are described below in more detail. H i holes and expanding shells. The Hα image in Figure 1 reveals the prominent superbubble Shell 1 and the fainter superbubble Shell 2 in N44. Shell 1, around the OB association LH47, has a size of 70 pc × 50 pc; while Shell 2, with no obvious OB association inside, is 60 pc across. The H i channel maps at Vhel = 292 -- 302 km s−1 (Figure 3) show clear H i holes at the central cavities of Shell 1 and Shell 2. Projected within the H i holes are high-velocity H i clumps (∼15 pc diameter) at 5h22m45s, −67◦56′00′′ (J2000) at the velocity range of Vhel ∼ 228 -- 245 km s−1(see Figure 3). These high-velocity H i clumps are most likely the brightest parts of the approaching hemisphere of an expanding H i shell associated with the optical shell. The receding hemispheres of these shells expand more slowly. The expansion of these H i shells is better visualized in the position -- velocity plots (or L − V diagrams) in Figure 4. These shells will be analyzed in detail in §4. Besides the H i shells with obvious optical counterparts, we see two H i holes, centered at Vhel = 297 km s−1, without known expanding shells in the optical. Inspected closely, the H i hole at 5h22m00s, −67◦51′50′′ (J2000) is associated with sharp, curved optical filaments that appear to delineate a shell structure, while the H i hole at 5h21m07s, −67◦51′45′′ (J2000) appears to be associated with only faint, irregular optical filaments. Interestingly, as shown in Figure 5, both of these H i holes encompass regions of bright diffuse X-ray emission reported by Chu et al. (1993). The relationship between the X-ray-emitting ionized gas and the H i gas will be discussed further in §6. Acceleration of H i. The H i channel maps in Figure 3 show that the bulk of H i gas is detected in the velocity range of 295 -- 305 km s−1. High-velocity H i gas with ∆V up to −80 km s−1 is detected in a portion of the SNR 0523−679. This high-velocity gas must have been accelerated by the SNR. High-velocity H i gas (Vhel = 233 km s−1) is also detected in the region between the SNR and the superbubble Shell 1. This gas might consist of two components, one accelerated by the SNR and the other by Shell 1, as the H i contours have an excellent correspondence with an Hα blister on the periphery of Shell 1 . -- 6 -- Compression of H i. In the channel maps near the bulk velocity, H i peaks are seen in regions between expanding H i shells. It is possible that these peaks represent compressions caused by two shells expanding into each other. One example is the region at 5h22m40s, −67◦54′40′′ (J2000) between the SNR 0523−679 and the superbubble Shell 1; the H i column density, N(H i) = 3.9 × 1021 cm−2, is a factor of 2 higher than those of adjacent regions along the rim of shell.6 Another example is the region at 5h23m06s, 5h21m58s, −67◦55′00′′ (J2000) bounded by Shell 1, Shell 2, and the H i hole to the north. The H i column density, N(H i) = 3.5 × 1021 cm−2, is a factor of 2 higher than those of adjacent regions along the rim of shell. It is unlikely that the H i peaks are caused by a low ionization in these directions, since the ionizing source of Shell 1 is located at the center of the shell. These H i peaks must be real enhanced column densities as a result of compression between expanding shells. Opening of a Breakout in H i. A breakout in Shell 1 toward the south has been suggested by Chu et al. (1993), based on an extension of diffuse X-ray emission toward the group of H ii regions associated with the OB association LH49. This explanation has been confirmed by the presence of high-velocity ionized gas and a lower plasma temperature in the breakout region (Magnier et al. 1996). The H i channel maps at Vhel of 300 -- 305 km s−1 show a clear extension of the H i hole to the southeast, coinciding with the X-ray extension. This extension of the H i hole is likely the opening through which the breakout occurs. Near the southeast end of the X-ray extension, we see an H i clump peaking at Vhel = 285 km s−1 with a column density N(H i) of 2.8 ×1021 cm−2, a factor of 3 higher than the surrounding region. This H i clump might have been impinged upon and compressed by the flow of hot, ionized gas in the breakout. 4. The Superbubble Structure of Shell 1 and Shell 2 Previous studies of the superbubble structure of N44 have used observations of only ionized gas (e.g., Chu et al. 1993; Magnier et al. 1996). It is found that the observed kinetic energy of the ionized Shell 1 of N44 is at least an order of magnitude lower than the value expected in Weaver et al.'s (1977) superbubble model (Magnier et al. 1996). Our new H i observation of N44 reveals the existence of a neutral gas associated with Shell 1 and possibly Shell 2 as well. The kinetic energy of the neutral shell may alleviate the aforementioned discrepancy in energies. Below we derive the dynamic parameters separately for the ionized and neutral components of Shell 1. The energetics of Shell 2 is discussed near the end of this section. 6The H i velocity profile of this region has an extended red wing up to Vhel ∼ 320 km s−1. However, the channel map centered at 310 km s−1 shows an unrelated feature extending from this region to the northeast. The compression is referred to the dense clump at the systemic bulk velocity. -- 7 -- The expansion of Shell 1 is not uniform. Its systemic velocity, determined from the Hα velocities at the rim of the shell, is 295 -- 298 km s−1. Referenced to this systemic velocity, the receding hemisphere of the ionized gas shell has an expansion velocity of 30 km s−1, and the approaching side 45 km s−1. We will analyze the physical parameters of the approaching and receding hemispheres separately. These parameters are summarized in Table 2. The Hα echelle observation of the central region of Shell 1 shows that the receding component is three times as bright as the approaching component. Using photoelectrically calibrated PDS scans of the Curtis Schmidt plates of Kennicutt & Hodge (1986) and together with our CCD Hα image, we derive an emission measure of 214 cm−6 pc for the receding component and 71 cm−6 pc for the approaching component. To derive the rms shell density, we have assumed a uniform thin shell for Shell 1 to estimate its shell thickness. The Hα surface brightness profile of Shell 1 shows a peak-to-center ratio of 13, indicating a fractional shell thickness ∆R/R of 0.02. The FWHM of the brightness peak at the shell rim is 0.1 times the shell radius. Given the nonuniform densities, the fractional shell thickness is most likely within the range of 0.02 -- 0.1. We have derived the rms density, mass, and kinetic energy of the shell for shell thicknesses of 0.02 and 0.1, respectively, and listed them in Table 2. The rms density in the receding hemisphere is almost twice as high as that in the approaching hemisphere, but the kinetic energy of the receding hemisphere is 21% lower than that of the approaching hemisphere. The total mass and kinetic energy of the H ii shell are 2580 M⊙ and 3.5 × 1049 ergs for a ∆R/R of 0.1, or 1210 M⊙ and 1.6 × 1049 ergs for a ∆R/R of 0.02. The H i material associated with Shell 1 of N44 shows a typical expansion structure in the position -- velocity diagrams (Figure 4) along cuts crossing the central portion of Shell 1. H i is detected in the velocity range of Vhel ∼ 220 -- 327 km s−1. The velocity gradient of the high-velocity component in the central cuts of Shell 1 indicates an expansion motion. At the center of Shell 1, the receding side of the H i shell is expanding at a lower velocity, 28±5 km s−1, than the approaching side, 50±5 km s−1. These expansion velocities are comparable to those of the ionized component of Shell 1. Following the analysis of the ionized shell, we discuss the receding and approaching hemispheres of the H i shell separately. The column densities, masses, and kinetic energies are listed in Table 2. The H i column density of the receding hemisphere is 2.5 times that of the approaching hemisphere, but the kinetic energy of the receding hemisphere is 21% lower. This contrast is remarkably similar to that in the H ii shell. The total mass of the H i shell is 8640 M⊙, and the total kinetic energy in H i is 1.1 × 1050 ergs. Note that the total mass and total kinetic energy of the H i component of Shell 1 are 3 -- 7 times higher than those of the H ii component. This implies that analyses of superbubble dynamics must include the neutral gas component! Note also that the asymmetric expansion of Shell 1, shown in both H i and H ii components, indicates a stratified interstellar medium with densities decreasing toward us. -- 8 -- We may re-examine the energetics of Shell 1 with the additional H i information. The total stellar wind and supernova energy input implied by the observed stellar content (Oey & Massey 1995), the total kinetic energy derived in this paper, and the thermal energy of the hot interior derived from X-ray observations (Magnier et al. 1996) are summarized in Table 2. In a pressure-driven superbubble, the thermal and kinetic energies are expected to be 35/77 and 15/77 times the total mechanic energy input from stellar winds and supernovae (Weaver et al. 1977; Mac Low & McCray 1988). The expected thermal energy and shell kinetic energy are also listed in Table 2. The observed shell kinetic energy is dominated by that of the expanding H i shell; still, there is at least a factor of 5 discrepancy between the observed and expected kinetic energies. The discrepancy might be caused by the breakout of Shell 1, since a large fraction of the stellar energy input may have been lost in the breakout. Other possibilities are discussed in §6.2. The superbubble Shell 2 of N44 appears as an H i hole in the channel maps as well as in the position-velocity diagrams (Figure 4). The receding side of the shell is detected along some cuts across the shell. The approaching side is much fainter; only a bright, high-velocity H i knot is detected. (The rest of the approaching hemisphere is probably below our detection limit.) In a similar way, we describe the physical parameters of the H i component of Shell 2. The receding part of the H i shell has an expansion velocity of 30±5 km s−1, lower than that of the approaching side, 50±5 km s−1. The asymmetric expansion of the H i in Shell 2 is similar to that of Shell 1; therefore, Shell 1 and Shell 2 probably share the same interstellar environment. The H i mass of Shell 2 is 5850 M⊙ in the receding hemisphere and 3090 M⊙ in the approaching hemisphere. The total kinetic energy of the H i in Shell 2 is 1.3 ×1050 ergs, 20% higher than that of the H i in Shell 1. The stellar content of Shell 2 has not been as well studied as that of Shell 1, and the X-ray observations of Shell 2 have much lower S/N ratios, hence Shell 2 cannot be modeled in as much detail as Shell 1. 5. The Physical Structure of the SNR 0523−679 The SNR 0523−679, centered at 5h23m06s, −67◦53′15′′ (J2000), was first diagnosed by its diffuse X-ray emission (Chu et al. 1993). The diffuse X-ray emission is surrounded by curved Hα filaments that are suggestive of a shell with a diameter of ∼3.′8. The Hα shell, called Shell 3 by Chu et al. (1993), is bounded by H ii regions on the west and south sides. The bright arc on the west side of the shell is also on the surface of the H ii region around the OB association LH48, indicating a physical interaction between the SNR and the H ii region. The southern part of the SNR shell is blended with a diffuse ring-like H ii region; however, it is not clear whether they physically interact. Within the boundary of the SNR shell, long Hα filaments and dust lanes run from the NE corner to the SW corner, dissecting the shell into two lobes. The X-ray emission shows a corresponding 2-lobe structure, which -- 9 -- could be caused by the absorption in the dust lane. The internal motion of the SNR is revealed in the echellograms at slit positions S4 and S5 (Figure 1). The velocities at the shell rim sampled by these two slits are Vhel = 297 -- 304 km s−1, which will be adopted as the systemic velocity of the ionized SNR shell. The echellograms show very different velocity structures in the north lobe and the south lobe of the SNR. The north lobe shows prominent, blue-shifted material in the Hα line; the brighter parts have velocity offsets up to −125 km s−1, while the fainter parts have velocity offsets up to −150 km s−1. The receding side of the north lobe has much smaller velocity offsets, ∼20 km s−1. The south lobe shows mostly red-shifted material in the Hα line, with velocity offsets up to +100 km s−1; however, blue-shifted material is also detected at the northern end of the south lobe, with velocity offsets up to −100 km s−1. For such an irregular pattern of expansion, it is impossible to determine the kinetic energy of the ionized gas in SNR 0523−679 using our limited echelle observations. We can nevertheless assume a uniform shell and derive upper limits on the mass and kinetic energy. The brightest part of the SNR along slit S5 has an emission measure of 25 cm−6 pc. The apparent width of the SNR shell, 6.′′5 or 1.6 pc, provides an upper limit on the real shell thickness. The corresponding rms ne is ∼ 4 cm−3. The upper limits on the mass and kinetic energy are thus 370 M⊙ and 6 × 1049 ergs. The H i gas associated with SNR 0523−679 has a very different velocity structure from that of the ionized gas described above. The bulk velocity of H i gas toward the SNR is Vhel = 306±5 km s−1. Since this velocity is similar to the SNR's systemic velocity determined with the Hα line, we will adopt this velocity as the systemic velocity of the H i gas associated with the SNR. The channel maps in Figure 3 show an H i hole at Vhel = 297 km s−1, although the size of the H i hole is smaller than that of the Hα shell. The receding side of the H i gas associated with the SNR is clearly seen in the channel map at Vhel = 325 km s−1, suggesting an expansion velocity of 20 -- 30 km s−1. The approaching side of the H i gas is more complex. Blue-shifted H i gas in the velocity range Vhel = 230 -- 245 km s−1 is seen toward the SNR. However, this H i gas also extends further south than the boundary of the SNR; the southern extension is particularly pronounced in the region between the SNR and Shell 1 at Vhel = 228 -- 235 km s−1. As we explained in §3, Shell 1 might be responsible for the acceleration of the high-velocity H i in the region between Shell 1 and the SNR. Interestingly, the highest-velocity (Vhel = 226 km s−1) H i component unambiguously associated with the SNR is located at the NE edge, instead of the center, of the optical shell. This might indicate that a breakout similar to that seen at the south edge of Shell 1 has occurred at this position. We may estimate the mass and kinetic energy of the H i gas associated with the SNR. The high-velocity H i gas is more extended than the optical boundary of the SNR. To exclude the contribution from other sources, we use only the part of H i gas projected -- 10 -- within the optical boundary of the SNR. The mass of the approaching side of the H i gas is 7460±1000 M⊙ and the kinetic energy of the SNR is 2.7 × 1050 ergs. The mass and kinetic energy of the H i gas are much higher than those of the ionized gas in the SNR. This is similar to what we have found for the superbubble Shell 1. The thermal energy of the SNR can be derived from the ROSAT and ASCA observations in X-rays. The best Raymond and Smith (1977) model fits of these two sets of data give very different plasma temperature and absorption column density; however, the thermal energies derived from these two sets of parameters are similar. The best fit to the ROSAT Position Sensitive Proportional Counter data (Chu et al. 1993) gives a plasma temperature of kT = 0.835 keV, an absorption column density of log NH = 20.68, and a normalization factor of log N 2 e V f /4πD2 = 10.817, in the cgs units, where Ne is the electron density, V is the volume, f is the volume filling factor, and D is the distance. The thermal energy of the SNR calculated from this model fit is 1.2 × 1050(f /0.5)1/2 ergs. The best fit to the ASCA Solid-state Imaging Spectrometers (SIS) data gives kT = 0.35 keV, log e V f /4πD2 = 11.326. The thermal energy from this model fit is 9.0 NH = 21.48, and log N 2 × 1049(f /0.5)1/2 ergs. Considering the uncertainties of filling factor and spectral fits, the thermal energy of the SNR is probably (1.0±0.5)× 1050 ergs. The thermal energy in SNR 0523−679 is similar to that in each of the shells in the SNR DEM L 316 (Williams et al. 1997), despite the fact that SNR 0523−679 is in a much more inhomogeneous medium and has a much more nonuniform expansion. The kinetic energy in the ionized gas of SNR 0523−679 is of a similar order of magnitude as the thermal energy of SNR 0523−679. This behavior is similar to what is observed in DEM L 316. The total kinetic energy (H i + ionized) of SNR 0523−679 is much larger than the thermal energy in the SNR's hot interior. This behavior is similar to what is seen in N44's superbubble Shell 1. 6. Discussion 6.1. Implications on the H i Structure of the LMC The ATCA H i survey of the LMC (Kim & Staveley-Smith 1997) shows H i holes with sizes ranging from a few tens of pc to ∼1400 pc. It is conceivable that these H i holes have been created by the combined effects of fast stellar winds and supernova explosions. However, most of these holes do not have any recognizable stellar content or corresponding ionized gas shells, making follow-up studies difficult, if not impossible. N44, containing three OB associations at three different evolutionary stages, offers an ideal site for us to study how massive stars shape the interstellar medium. The current H i structure in N44 is readily visible, whereas the ionized gas structure allows us to foresee the future H i structure after the ionizing fluxes have terminated at the demise of all massive stars in N44. From -- 11 -- the study of N44, we hope to gain insight into the general H i structure of the LMC. We first examine the validity of the apparent "holes" in the H i maps of N44. The integrated H i map of N44 shows two deep depressions. Both depressions have associated H i shells, and both shells have optical counterparts, the superbubbles Shell 1 and Shell 2. For Shell 1, the H i shell and the ionized gas shell have similar expansion patterns and velocities: both have the far side receding at 30 km s−1 and the near side approaching at 45 -- 50 km s−1. For Shell 2, the H i shell appears to expand faster than the ionized gas shell: the H i shell has expansion velocities of 30 -- 50 km s−1, while the ionized gas shell show only a velocity variation from 304 to 320 km s−1 (Meaburn & Laspias 1991). This different behavior might be caused by the different distribution of ionizing sources. Shell 1 has a central ionizing source, the OB association LH47, hence its neutral H i shell is adjacent and exterior to its ionized gas shell. Shell 2, on the other hand, has no central OB association, and its ionizing flux might be provided by LH47's O stars to the east of Shell 2. The structure of the ionized gas in Shell 2 thus depends of the relative locations of the ionizing stars, and would not have a one-to-one correspondence with the neutral H i gas shell. The channel maps of H i in N44 also show numerous holes. The most obvious H i holes are present in the map centered at the bulk velocity of N44, Vhel = 297 km s−1. In addition to the aforementioned superbubbles Shell 1 and Shell 2, two other holes are present to the north. As described in §3, one of these H i holes has Hα filaments delineating a shell structure, while the other has no evidence of a coherent shell structure in Hα images. These two H i holes must be real holes, as both regions show enhanced X-ray emission (see Figure 5), indicating that the holes are filled with 106 K hot gas. It might be argued that the enhanced X-ray emission is resultant from a smaller foreground absorption. Since the integrated H i map does not show obvious anti-correlation between H i column density and X-ray surface brightness in N44, we consider this argument unlikely. These apparent H i holes in the channel maps must correspond to real cavities. Conversely, we next examine the H i maps for holes at regions where we expect them. The SNR 0523−679, having a diameter of ∼ 4′, is well resolved by our H i maps. However, only a very shallow depression in H i column density is present in the integrated H i map or the channel map centered at 297 km s−1. This is perhaps understandable because the SNR 0523−679 is in a low density medium, indicated by its fast expansion and large size. The H i gas that has been accelerated by the SNR also shows a patchy structure, consistent with a low density medium around the SNR. We may also expect H ii regions to show up as holes in H i maps. However, neither the H ii region around LH48 nor the H ii region around LH49 shows a hole in the integrated or channel maps of H i. The lack of correspondence between young H ii regions and H i distribution indicates that the star formation must have taken place in molecular clouds and the H ii regions contain the dissociated and ionized molecular gas. This conclusion has been previous reached by Allen, Atherton, & Tilanus (1986) based on their H i observations -- 12 -- of the spiral galaxy M83. The presence of molecular cloud in N44 is evidenced in the CO map by Cohen et al. (1988). CO maps made with the ESO-SEST at a higher resolution show molecular material associated with the B, C, and D components of N44, the compact H ii regions along the west (the B component) and southwest (the C component) rim of Shell 1 and to the southeast (the D component, associated with LH 49) of Shell 1 (Israel et al. 1993; Chin et al. 1997). This molecular material must be the remnants of the natal clouds. Besides H i holes and expanding shells, we see clear evidence of acceleration of H i gas by superbubbles or SNRs, compression of H i gas between expanding shells, and breakout structures in N44. Similar phenomena must exist in other active star forming regions. It is conceivable that the interstellar medium has been shaped by multiple episodes of massive star formation, and that the complex H i structure represents the cumulative effects of the previous massive star formation. 6.2. H i Shell and Superbubble Dynamics The most remarkable lesson we have learned from the H i study of superbubble in N44 is that the neutral gas actually contains more mass and kinetic energy than the ionized gas! For example, in Shell 1 of N44, the mass and kinetic energy in the expanding H i shell is 3 -- 7 times those of the ionized gas shell. Thus, the H i gas is an important component in superbubbles. Previously, superbubble dynamics has been observed with optical emission lines, hence only the ionized gas shells have been considered. This clearly needs to be improved. It has been a standing problem that the observed kinetic energy in a wind-blown bubble is much too low compared to that expected from Weaver et al.'s (1977) pressure-driven bubble model (Oey & Massey 1995). We have summed the kinetic energy in the ionized gas shell and in the H i shell for the superbubble Shell 1 in N44, and found that the observed total kinetic energy is still at least a factor of 5 lower than expected. The observed ratio of kinetic energy to thermal energy is nevertheless closer to the theoretical value of 3/7, especially if the thermal energy in the breakout region is included. The superbubble dynamics of Shell 1 may have been significantly altered by the energy loss through the breakout; however, the breakout may not be the only cause for the discrepancy between the observed superbubble dynamics and theoretical predictions. The fact that the age of the OB association LH47, ≥ 10 Myr (Oey & Massey 1995), is much larger than the dynamic age of Shell 1, < 1 Myr, indicates that the formation of a superbubble did not start as soon as the central OB association was formed. It is possible that LH47 was formed in a molecular cloud, and the currently visible superbubble Shell 1 did not start to expand rapidly until it has broken out of the molecular cloud. This -- 13 -- hypothesis is supported by the remnant molecular clouds observed toward N44 (Israel et al. 1993; Chin et al. 1997). Only the youngest H ii regions are still associated with molecular material, and these youngest H ii regions show neither H i holes nor H ii shell structures. Similar explanation has been proposed for the superbubble in N11, and may be common among all superbubbles (Mac Low et al. 1997). Of course, off-center SNR shocks could have provided acceleration and contributes to the discrepancy between the observed superbubble dynamics and the theoretical predictions. Evidence of SNR shocks in N44 includes the excess X-ray emission (Chu et al. 1993) and strong [S II]/Hα ratios (Hunter 1994). Future observational studies of superbubble dynamics need to include both the neutral and ionized gas components and extend to superbubbles with a wide range of X-ray luminosities, so that additional SNR acceleration effects can be differentiated from normal superbubble dynamics. We thank the team members of the HI survey project -- R. Sault, K. C. Freeman, M. A. Dopita, M. J. Kesteven, D. McConnell, and M. Bessell. SK would like to thank P. R. Wood for helpful comments to improve the paper, and J. Mould and MSSSO for financial supports to make the collaborative visit to Illinois possible. YHC acknowledges the NASA grants NAG 5-3246 (LTSA) and NAG 5-1900 (ROSAT). RCS acknowledges NSF grant AST-9530747 and the Dean B. McLaughlin Fund at the University of Michigan. -- 14 -- REFERENCES Allen, R. J., Atherton, P. D., & Tilanus, R. P. J. 1986, Nature, 319, 296 Chin, Y.-N., Henkel, C., Whiteoak, J.B., Millar, T.J., Hunt, M.R., & Lemme, C. 1997, A&A, 317, 548 Chu, Y.-H., & Mac Low, M.-M. 1990, ApJ, 365, 510 Chu, Y.-H., Mac Low, M.-M., Garc´ıa-Segura, G., Wakker, B., & Kennicutt, R. C., Jr. 1993, ApJ, 414, 213 Cohen, R. S., Dame, T. M., Garay, G., Montani, J., Rubio, M., & Thaddeus, P. 1988, ApJ, 331, L95 Henize, K. G. 1956, ApJS, 2, 315 Hunter, D. 1994, AJ, 107, 565 Israel, F.P., Johansson, L.E.B., Lequeux, J., Booth, R.S., Nyman, L.-A., Crane, P., Rubio, M., de Graauw, Th., Kutner, M.L., Gredel, R., Boulanger, F., Garay, G., & Westerlund, B. 1993, A&A, 276, 25 Kennicutt, R. C., & Hodge, P. W. 1986, ApJ, 306, 130 Kim, S., & Staveley-Smith, L. 1997, in Star Formation Near and Far (Maryland:AIP), in press Kim, S., Staveley-Smith, L., Sault, R., Freeman, K.C., Dopita, M.A., Kesteven, M.J., McConnell, D., & Bessell, M. 1997, in Preparation Lucke, P. B. & Hodge, P. W. 1970, AJ, 75, 171 Mac Low, M.-M., Chang, T. H., Chu, Y.-H., Points, S. D., Smith, R. C., and Wakker, B. P. 1997, ApJ, in press Mac Low, M.-M., & McCray, R. 1988, ApJ, 324, 776 Magnier, E. A., Chu, Y.-H., Points, S. D., Hwang, U., & Smith, R. C. 1996, ApJ, 464, 829 McKee, C. F., & Ostriker, J. P. 1977, ApJ, 218, 148 Meaburn, J. & Laspias, V. N. 1991, A&A, 245, 635 Norman, C. A., & Ikeuchi, S. 1989, ApJ, 345, 372 Oey, M. S. 1996, ApJ, 467, 666 Oey, M. S., & Massey, P. 1995, ApJ, 452, 216 Raymond, J. C., & Smith, B. W. 1977, ApJS, 35, 419 Sramek, R.A. & Schwab, F.R. 1989, in Synthesis Imaging in Radio Astronomy, eds. Perley, R.A., Schwab, F.R., & Bridle, A.H.(San Francisco:ASP), 123 Wang, Q., & Helfand, D.J. 1991, ApJ, 373, 497 -- 15 -- Weaver, R., McCray, R., Castor, J., Shapiro, P., Moore, R., 1977, ApJ, 218, 377 Williams, R. M., Chu, Y.-H., Dickel, J. R., Beyer, R., Petre, R., Smith, R. C., & Milne, D. K. 1997, ApJ, 618 This preprint was prepared with the AAS LATEX macros v4.0. -- 16 -- FIGURE CAPTIONS Fig. 1. -- Hα image of the N44 complex and echellograms. The echelle slit positions are marked in the Hα image. The slit 'S3' samples the center of the superbubble Shell 1. The slit "S1" samples the breakout region, or the X-ray South Bar (Magnier et al. 1996). 'S4' and 'S5' sample the lobes of the SNR 0523−679. The two narrow lines in the echellograms are the geocoronal Hα and the telluric OH line blend at 6577.284 A. Fig. 2. -- An ATCA integrated HI map of the N44 complex. The optically defined 'Shell 1' and 'Shell 2' are labeled on the HI contours. Corresponding HI holes in these positions can be clearly seen. Fig. 3. -- The ATCA channel maps of N44's HI datacube. The heliocentric velocity is given in the upper left corner. The H i contours are overplotted on (a) greay-scale images of H i themselves, and (b) an Hα image. The contour levels are 0.07, 0.10, 0.15, 0.20, 0.25, 0.30, 0.35, 0.40, 0.45 Jy/Beam. Fig. 4. -- The position-velocity plots of HI in the N44 complex. The contours corresponds to the associated pixel brightness ≥ 2.5 σ. The high-velocity gas associated with SNR might have unresolved gas component, which relate with an Hα blister on the periphery of 'Shell 1'. Fig. 5. -- H i contours overplotted on an X-ray image of N44. The H i contours are derived from the channel map centered at Vhel = 297 km s−1. The H i contour levels are 0.07, 0.10, 0.15, 0.20, 0.25, 0.30, 0.35, 0.40, 0.45 Jy/Beam. The X-ray image is a smoothed ROSAT PSPC image from Chu et al. (1993). -- 17 -- Table 1: Journal of Echelle Observations No. RA (J2000) Dec (J2000) Slit Exposure Remarks (hh mm ss) 05 22 32 05 22 32 05 22 17 05 23 01 05 23 05 05 22 32 1 2 3 4 5 6 ′ ( ◦ −68 00 06 −68 00 46 −67 56 18 −67 52 02 −67 54 09 −68 00 26 ′′ ) Orientation E -- W E -- W E -- W E -- W N -- S E -- W (s) 1200 1200 South breakout region 40′′S of slit 1 2×1200 Center of Shell 1 2×1200 2×1200 SNR SNR 20′′S of slit 1 1200 -- 18 -- Table 2: The Physical Parameters of Shell 1 and Shell 2 Parameter Shell 1a Shell 1b Shell 2 Front Back Front Back Front Back 2.3 5.6 950 45 1.9 -- 2.3 -- 9.7 1630 -- -- 30 1.5 -- -- -- -- -- -- 1.6 3090 50±5 7.7 3.0 5850 30±5 5.2 13 The H ii hemisphere: Shell Thickness (pc) ne (cm−3) Mass (M⊙) Vexp (km s−1) Ekin (1049 ergs) The H i hemisphere: N(HI) (1020 cm−2) Mass (M⊙) Vexp (km s−1) Ekin (1049 ergs) The whole shell: H ii Ekin (1049 ergs) H i Ekin (1049 ergs) Total Ekin, observed (1049 ergs) Eth, observed in shell (1049 ergs) Eth, observed in shell & breakout (1049 ergs) Wind and SN Energy Input (1049 ergs) Total Ekin, expected (1049 ergs) Eth, expected (1049 ergs) 0.5 11.9 440 45 0.89 3.6 6180 28 4.8 0.5 20.7 770 30 0.68 1.4 2460 50 - 60 6.1 1.6 -- 3.4 11 13 -- 14 3 -- 20 13 -- 30 400 -- 700 78 -- 136 200 -- 300 Notes: a with an ionized shell thickness ∆R/R of 0.02 b with an ionized shell thickness ∆R/R of 0.1
astro-ph/0102061
1
0102
2001-02-04T03:11:45
A probabilistic formulation for Empirical Population Synthesis: Sampling methods and tests
[ "astro-ph" ]
We present a probabilistic formulation of the classical problem of synthesizing spectral properties of a galaxy using a base of star clusters. The problem consists of estimating the population vector x, composed by the contributions of n_star base elements to the integrated spectrum of a galaxy, and the extinction A_V, given a set of absorption line equivalent widths and continuum colors. The formalism is applied to the n_star = 12 base defined by Schmidt etal and subsequently used in several studies. The 13-D parameter space is explored with a Markov chain Monte Carlo sampling scheme based on the Metropolis algorithm, which produces a smooth and efficient mapping of the P(x,A_V) probability distribution. This version of Empirical Population Synthesis is used to investigate the ability to recover the detailed history of star-formation and chemical evolution using this spectral base. This is studied as a function of (1) the magnitude of the measurement errors and (2) the set of observables used in the synthesis. Only for extremely high S/N all 12 base proportions can be accurately recovered, though the observables are very precisely reproduced for any S/N. Furthermore, the individual mean x components are biased in the sense that components which carry a large fraction of the light tend to share their contribution preferably among components of same age. This compensation effect is linked to noise-induced linear dependences in the base, which very effectively redistribute the likelihood in x-space. The age distribution, however, can be satisfactorily recovered for realistic data quality. (abridged)
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000–000 (0000) Printed 6 March 2018 (MN LATEX style file v1.4) A probabilistic formulation for Empirical Population Synthesis: Sampling methods and tests Roberto Cid Fernandes1,2⋆†, Laerte Sodr´e Jr.3‡, Henrique R. Schmitt4§, Joao R. S. Leao 2¶ 1Depart. of Physics & Astronomy, Johns Hopkins University, 3400 N. Charles St., Baltimore, MD 21218, USA 2Depto. de F´ısica - CFM - Universidade Federal de Santa Catarina, CP 476, Campus Universit´ario, Trindade, 88040-900, Florian´opolis, SC, Brazil 3Depto. de Astronomia, Instituto Astronomico e Geof´ısico - USP, Av. Miguel Stefano 4200, 04301-904 Sao Paulo, Brazil 4 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA 6 March 2018 ABSTRACT We revisit the classical problem of synthesizing spectral properties of a galaxy using a base of star clusters, approaching it from a probabilistic perspective. The problem consists of estimating the population vector x, composed by the contributions of n⋆ different base elements to the integrated spectrum of a galaxy, and the extinction AV , given a set of absorption line equivalent widths and continuum colors. The formalism is applied to the base of 12 elements defined by Schmidt et al. (1991) as correspond- ing to the principal components of the original base employed by Bica (1988), and subsequently used in several studies of the stellar populations of galaxies. The explo- ration of the 13-D parameter space is carried out with a Markov chain Monte Carlo sampling scheme based on the Metropolis algorithm. This produces a smoother and more efficient mapping of the P (x, AV ) probability distribution than the traditionally employed uniform-grid sampling. This new version of Empirical Population Synthesis is used to investigate the abil- ity to recover the detailed history of star formation and chemical evolution using this spectral base. This is studied as a function of (1) the magnitude of the measurement errors and (2) the set of observables used in the synthesis. Extensive simulations with test galaxies are used for this purpose. Emphasis is put on the comparison of input parameters and the mean x and AV associated with the P (x, AV ) distribution. It is found that only for extremely low errors (S/N > 300 at 5870 A) all 12 base propor- tions can be accurately recovered, though the observables are recovered very precisely for any S/N . Furthermore, the individual x components are biased in the sense that components which carry a large fraction of the light tend to share their contribution preferably among components of same age. Old, metal poor components can also be confused with younger, metal rich components due to the age × metallicity degener- acy. These compensation effects are linked to noise-induced linear dependences in the base, which very effectively redistribute the likelihood in x-space. The age distribution, however, can be satisfactorily recovered for realistic S/N (∼ 30). We also find that synthesizing equivalent widths and colors produces better focused results that those obtained synthesizing only equivalent widths, despite the inclusion of the extinction as an extra parameter. Key words: galaxies: evolution - galaxies: stellar content - galaxies: statistics ⋆ e-mail: [email protected] † Gemini fellow ‡ e-mail: [email protected] § e-mail: [email protected] ¶ e-mail: [email protected] R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao 2 1 INTRODUCTION The stellar populations of galaxies carry a record of their star-forming and chemical histories, from the epoch of for- mation to the present. Their study thus provide a powerful tool to explore the physics of galaxy formation and evolu- tion. Despite the detailed understanding of stellar evolution and of the properties of single stellar populations such as star clusters achieved over the past century, uncovering the history of galaxies through the study of their integrated light has proven to be a difficult problem (Pickles 1985; Worthey 1994 and references therein). Two global approaches have been developed to tackle this issue. The first of these is Evolutionary Population Syn- thesis, which performs ab initio calculations of the spectral evolution of galaxies on the basis of stellar evolution the- ory, stellar spectral libraries, and prescriptions for the ini- tial mass function, star formation rate and chemical evolu- tion (Tinsley 1972; Larson & Tinsley 1974, 1978; Guider- doni & Rocca-Volmerange 1987; Fioc & Rocca-Volmerange 1997; Charlot & Bruzual 1991; Bruzual & Charlot 1993; Lei- therer et al. 1996, 1999 and references therein). The other approach is Empirical Population Synthesis (EPS here- after), also known as 'Stellar Population Synthesis with a Database', which uses observed properties of stars or star clusters as a base on which to decompose the population mix in galaxy spectra (Spinrad & Taylor 1971; O'Connell 1976; Faber 1972; Pickles 1985; Bica 1988; Pelat 1997, 1998; Bois- son et al. 2000). Of particular interest is the EPS method of synthesizing the spectra of galaxies using a base of star clus- ters. Perhaps the most appealing property of this method is the fact that the results do not hinge on assumptions about stellar evolution tracks nor initial mass functions, as these are 'mother nature given' in the empirically built base spec- tra. This technique has been pioneered by the work of Bica (1988, hereafter B88), where the stellar content of a large sample of galaxies was synthesized using the spectral base of star clusters described in Bica & Alloin (1986a, 1986b, 1987). Since then, several studies made use of this technique (Schmidt, Bica & Dottori 1989; Bica, Alloin & Schmidt 1990; Jablonka, Alloin & Bica 1990; Schmidt, Bica & Alloin 1990; Jablonka & Alloin 1995; De Mello et al. 1995; Bonatto et al. 1996, 1998, 1999, 2000; Schmitt, Storchi-Bergmann & Cid Fernandes 1999; Kong & Cheng 1999; Raimann et al. 2000). In its original version (B88), this method made use of a large base of n⋆ = 35 elements, combined in different pro- portions (xi; i = 1 . . . 35) to synthesize a number of equiv- alent widths (W ) of conspicuous absorption features. An uniform sweep of the parameter space along paths consis- tent with simple predefined chemical evolution scenarios was performed, and all combinations which produced W 's within 10% of the observed values were considered 'solutions'. The final population vector x was taken to be the arithmetic mean of all such sampled solutions. One of the main concerns raised by B88 method is that of the uniqueness of the solution, as first discussed by Schmidt et al. (1991, hereafter S91). With a 35-D popula- tion vector and at most 9 W 's to be synthesized, one has a highly degenerate algebraic system. As shown by S91, how- ever, the original base was highly redundant, with several linearly dependent elements, and others which were practi- cally so, as deduced by a principal component analysis. S91 then proposed the use of a reduced base of 12 elements. S91 also criticized B88's uniform sampling of the x space (a 'dis- crete combination procedure' in their terms) and the use of a 'mean solution'. Instead, they developed a constrained min- imization procedure, which, as shown by their simulations with test galaxies built out of the base, is able to retrieve with good accuracy the correct population vector. Another difference of the S91 method compared to B88 is the possi- bility to search for solutions along the entire age × metal- licity plane. B88 assumed that the chemical enrichment of the galaxy occurred during its formation (∼ 10 Gyr ago), with all stars younger than that having the same Z as the oldest most metal rich component. Although this may be a reasonable approximation for the nuclear region of isolated galaxies, it does not allow for the effects of mergers, inflows or outflows, and is not applicable to high redshift sources, where the spectra usually integrate over the entire galaxy, including regions with different star formation and chemical histories (Jablonka et al. 1990; Jablonka & Alloin 1995). The next improvement in this technique was the intro- duction of continuum colors in the synthesis. This was done by Schmitt, Bica & Pastoriza (1996), who, as B88, use a discrete combination procedure search for acceptable solu- tions, sweeping the parameter space along chemical enrich- ment paths. This version of EPS has been used in several subsequent studies (Bonatto et al. 1996, 1998, 1999, 2000; Kong & Cheng 1999). Another step was taken by Bonatto, Bica & Alloin (1995), who extended the star cluster base to the space ultraviolet (1200–3200 A). They too use an arith- metic average to represent the solution of the synthesis. In two recent papers, Pelat revisited the EPS problem of synthesizing W 's using an elegant formalism based on convex algebra concepts. In Pelat (1998) he shows how the algebraic solution of the EPS problem can be narrowed down to a sub-space of x delimited by a set of 'extreme-solutions', any convex combination of which yields an exact solution in the underdetermined cases (more parameters than con- straints). His results shed new light into the issue of alge- braic degeneracy, which has always haunted EPS due to the widespread belief that with more parameters than observ- ables one is bound to fit the data in one way or another. He proposes a quick test of whether the W 's of a galaxy are synthesizable by computing the region in W -space spanned by all physical combinations of the base elements. We veri- fied that in the case of the 12 elements base used by S91 this region is narrow (Leao 2001), such that small measurement errors are enough to place objects outside this 'synthetic domain', hence preventing the existence of mathematically exact solutions. In real applications, one therefore expects the degeneracy of EPS to be more of a statistical nature than algebraic. In these cases, as well as in overdetermined problems, Pelat (1997) demonstrates that the best model is to be found along the boundary of the synthetic domain by minimizing an appropriately defined χ2-like distance. Un- certainties in the population vector can also be readily com- puted, as shown by Moultaka & Pelat (2000). Boisson et al. (2000) recently applied Pelat's method to a sample of 12 active galaxies from Serote Roos et al. (1998). They synthesize nW = 47 absorption lines with a base of n⋆ = 30 stars from the stellar spectral libraries of Serote Roos, Boisson & Joly (1996) and Silva & Cornell (1992). Their results nicely illustrate the power of this new EPS Empirical Population Synthesis: Sampling methods and tests 3 technique, which will certainly be extended to larger sam- ples and other classes of galaxies in future studies. The B88 method and its variants, on the other hand, already have a large body of literature associated to it. Nevertheless, this popular and attractive EPS technique, while intuitively ac- ceptable, has neither been adequately tested nor cast onto a sound mathematical formalism which supports its applica- tion. Furthermore, it gives no measure of the uncertainties or potential biases on the population vector, thus giving no simple way to prevent overinterpretations of the resulting population of the age × metallicity plane. These issues are the focus of the present paper. Our main goals are to: (i) Elaborate a mathematically consistent formulation of the EPS problem in the context of probability theory. (ii) Develop and test a sampling method to explore the pa- rameter space in EPS problems. In particular, we aim to improve upon the Schmitt et al. (1996) technique of syn- thesizing both equivalent widths and colors, but within the context of a probabilistic formulation. (iii) Use the method to investigate the effects of measure- ment errors and of the specific set of observables used upon the results of the synthesis with the reduced B88 star clus- ter base described in S91. In particular, we wish to evaluate whether these factors introduce biases in the population mix inferred from this popular synthesis technique and at what level of detail can its results be trusted. This paper is organized as follows: Item (i) of the above list is addressed in section 2. In section 3 we deal with point (ii). This is done by investigating the applicability of a Metropolis algorithm to sample the parameters probability distribution in EPS, and testing the method with simula- tions based on Bica's spectral base. In section 4 we address item (iii) by means of extensive simulations which explore the ability to recover the detailed star-formation and chem- ical histories of galaxies with this base. Finally, section 5 summarizes our main results. 2 FORMALISM 2.1 Basic equations EPS studies seek to find combinations of a spectral base which reproduce a given set of measured observables, of- ten taken as the equivalent widths Wj of nW conspicuous absorption features. The base consists of n⋆ elements repre- senting well defined simple stellar populations such as star- clusters, with equivalent widths W ⋆ ij and corresponding con- tinuum fluxes over the lines F ⋆ ij (j = 1, . . . nW , i = 1, . . . n⋆) normalized at a reference wavelength. Denoting by xi the fractional contribution of the i-th base element to the total flux at the reference wavelength, one obtains a system of nW equations ; j = 1, ..., nW (1) Wj = Wj(x) = Pn⋆ i=1 W ⋆ Pn⋆ i=1 F ⋆ ijF ⋆ ijxi ij xi for the synthetic W 's (e.g., Joly 1974). This set of constraints can be augmented by synthesizing nC observed continuum fluxes Ck (k = 1, . . . nC ), provided allowance is made for reddening, here parametrized by the V-band extinction AV . Assuming all n⋆ populations are equally reddened, one ob- tains Ck = Ck(x, AV ) = gk(AV ) n⋆ Xi=1 C ⋆ ikxi; k = 1, ..., nC (2) where a distinction is made between C ⋆ and F ⋆, since the nC fluxes to be synthesized need not correspond to the wave- lengths of the nW absorption lines. We hereafter refer to the Ck's as continuum colors, as they are in fact ratios of contin- uum fluxes with respect to the continuum at the normaliza- tion wavelength. The function gk(AV ) reddens the normal- ized color at wavelength λk by Aλk according to a specified reddening law. Finally, physical solutions must further satisfy the nor- malization and positivity constraints: n⋆ Xi=1 xi = 1; xi ≥ 0 for i = 1, ..., n⋆, and AV ≥ 0. (3) The normalization condition effectively reduces one de- gree of freedom. When modeling colors, however, one has to introduce AV as a further parameter, so that the number of parameters still is n⋆. 2.2 Probabilistic formulation The data D to be modeled are thus composed of a set of nobs = nW + nC observables. The measurement errors in each observable, collectively denoted by σ, are assumed to be known. Given these, the problem of EPS is to estimate the population vector x and the extinction AV that 'best' represents the data according to a well defined probabilistic model. It is equally important to estimate the uncertainties in the parameters, as these prevent over-interpreting the re- sulting mixture of stellar populations at a level of detail not warranted by the data or by intrinsic limitations of the base. The probability of a solution (x, AV ) given the data D and the errors σ, is given by Bayes theorem: P (x, AV D, σ, H) = P (Dx, AV , H)P (x, AV σ, H) P (DH) . (4) H summarizes the set of assumptions on which the in- ference is to be made. They include: the mapping between parameters and observables is given by eqs. (1) and (2); x and AV must satisfy the constraints expressed in eq. (3); the stellar population in the target galaxy is well represented by the base elements; the observational errors are Gaussian; the reddening law is known. The likelihood P (Dx, AV , σ, H) is a measure of how good (or bad) is the fitting to the data for model parameters x and AV . Under the hypothesis of Gaussian errors, the probability of the data given the parameters is P (Dx, AV , σ, H) ∝ e−E , with E defined as half the value of χ2: E (x, AV ) = χ2(x, AV ) = 1 2 1 2 nW Xj=1 j − Wj(x) σ(Wj) (cid:18) W obs (cid:19)2 + 1 2 nC Xk=1 (cid:18) C obs k − Ck(x, AV ) σ(Ck) (5) (cid:19)2 (6) 4 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao The likelihood, as the total χ2, separates into a W and a color related term, of which only the latter depends on AV . The normalizing constant P (DH) in Bayes theorem (the 'evidence of H') is irrelevant to the level of inference discussed here, i.e., the estimation of x and AV . It can be used, for instance, to compare different spectral bases. P (x, AV H) is the joint a priori probability distribution of x and AV , and states what values the model parame- ters might plausibly take. For instance, physically accept- able population fractions should have priors that are zero in regions of the parameter space where the constraints of positivity and unit sum are not satisfied. A prior that is uni- form on the parameters and includes these constraints is a non-informative prior, because it does not impose any con- straints on the solutions besides those expressed by eq. (3). This is the prior that will be used in this work. For a non-informative prior the posterior probability P (x, AV D, σ, H) is simply proportional to the likelihood: P (x, AV D, σ, H) ∝ e−E(x,AV ) (7) This expression contains the full solution of the EPS problem, as embedded in it is not only the most proba- ble model parameters but also their full probability distri- butions. Furthermore, projected posterior distributions for any of the n⋆ parameters can be obtained by marginalizing eq. (7) with respect to all other parameters. For instance, the probability density of proportion xi of the i-th base element is P (xiD, σ, H) = Z . . .Z P (x, AV D, σ, H)dx1 . . . (8) dxi−1dxi+1 . . . dxn⋆dAV and equivalently for AV . Similarly, one can construct joint posteriors for, say, the total proportion of all base compo- nents of same age or Z, or for the mean age and Z of the stellar population. This would of course provide a coarser description of a galaxy's stellar content than the individual posteriors P (xiD, σ, H), but that may be all that is possible under some circumstances, such as when only a reduced set of observables is available or when observational errors are large (section 4). It is easy to see how the B88 and Schmitt et al. (1996) synthesis techniques fit into this general probabilistic for- mulation. By synthesizing the observables within ∼ 10% 'error boxes', these authors implicitly assumed a box-car likelihood function P (Dx, AV , σ, H), whereas by perform- ing arithmetic means over all accepted (x, AV ) combina- tions in a uniform grid they are implicitly sampling the cor- responding posterior probability distributions. Also, the a priori constraints on the occupation of the age × Z plane imposed by these authors (but relaxed in subsequent works) simply reflect their use of an informative prior. This gives some formal justification for their heuristically designed EPS method. It is therefore reasonable to expect that an imple- mentation of EPS based on the formalism presented here should give results roughly compatible with those previously obtained. An uniform sweep of the (x, AV ) space is however a very inefficient sampler for (8), so some improvement is needed there. This is discussed next. 3 SAMPLING METHOD AND TESTS Despite its formal merits, at the computational level our probabilistic approach faces the same basic difficulty as pre- vious EPS codes, namely, the high dimensionality of the problem. For spectral bases with astrophysically interesting resolution in age and metallicity the number of elements n⋆ quickly becomes large enough to render the exploration of the parameter space a non-trivial task. Before discussing methods to sample the (x, AV ) space, we present the spec- tral base used in this work (§3.1) and describe how we deal with the uncertainties in the observables (§3.2). 3.1 The spectral base and observables The base used for the synthesis here is that of S91. It con- tains n⋆ = 12 population groups, spanning five age bins-10 Gyr (which actually represent globular cluster-like popula- tions), 1 Gyr, 100 Myr, 10 Myr and HII (corresponding to current star formation, and represented by a pure Fλ ∝ λ−2 continuum based on the spectrum of 30 Dor)-and four metallicities-0.01, 0.1, 1 and 4 Z⊙ (Table 1). The observ- ables in this base comprise the nW = 9 equivalent widths of the absorption lines CaII K λ3933, CN λ4200, G band λ4301, MgI λ5175, CaII λ8543, CaII λ8662, Hδ, Hγ and Hβ, as well as nC = 7 continuum fluxes at selected pivot wavelengths: 3290, 3660, 4020, 4510, 6630, 7520 and 8700 A, all normal- ized to 5870 A. According to Bica & Alloin (1986a, 1987) and Bica, Alloin & Schmitt (1994), the equivalent width windows and continuum points were defined based on very high signal to noise spectra of galaxies, with the express goal of using them to synthesize the stellar population of galax- ies. These spectra were obtained by creating two average spectra, of blue and red galaxies, where the stellar popula- tion features can be clearly traced. The base values for these quantities have been previously published in Bica & Alloin (1986b, 1987) and Schmitt et al. (1996), and some were mea- sured from data in Bica et al. (1994). These are recompiled in Table 2 , as they bear a direct impact on the analysis below. There are thus a maximum of 9 + 7 = 16 observ- ables to model with n⋆ = 12 parameters: n⋆ − 1 population fractions plus AV . However, in several of the tests presented below we shall make use of a reduced subset of observables, as observational data sets seldom cover the whole spectral range spanned by this base. According to Bica & Alloin (1986a, 1987) and Bica, Al- loin & Schmitt (1994), since the ultimate goal of the contin- uum points and W 's is to synthesize the stellar population of galaxies, the W windows and continuum points were de- fined based on very high signal to noise spectra of galaxies. These spectra were obtained by creating two average spec- tra, of blue and red galaxies, where the stellar population featuress can be identified. The reddening of the colors is modeled with the extinc- tion law described in Cardelli, Clayton & Mathis (1989, with RV = 3.1). 3.2 Errors in the observables The measurement errors in the observables play a key role in determining the structure of the likelihood function. Whereas such errors are available when analyzing observed Empirical Population Synthesis: Sampling methods and tests 5 Base Elements Used HII 10 Myr 100 Myr 1 Gyr 10 Gyr log(Z/Z⊙) 12 10 11 8 9 5 6 7 1 2 3 4 0.6 0.0 -1.0 -2.0 Table 1. Ages, metallicities and numbering convention for the star clusters in the base. # 1 2 3 4 5 6 7 8 9 10 11 12 1 2 3 4 5 6 7 8 9 10 11 12 Equivalent Widths (A) K CN G MgI CaT1 CaT2 Hδ 21.1 17.3 11.0 4.7 17.3 14.0 8.9 4.5 3.8 2.6 2.2 0.0 0.34 0.48 0.72 0.94 1.03 1.03 1.03 1.89 1.89 1.55 1.55 2.34 17.5 12.0 4.5 0.5 13.9 9.6 3.6 3.0 2.2 1.4 1.1 0.0 0.48 0.59 0.78 0.96 1.05 1.05 1.05 1.75 1.75 1.34 1.34 2.03 11.1 9.3 5.9 2.5 9.0 7.4 4.6 1.5 1.2 0.3 0.3 0.0 10.1 7.4 3.8 0.7 9.1 6.2 3.2 3.2 2.5 2.5 2.0 0.0 6.8 5.5 3.4 1.2 6.8 5.5 3.4 6.8 5.5 8.2 6.6 0.0 6.0 5.0 3.3 1.6 6.0 5.0 3.3 6.0 5.0 6.9 5.8 0.0 Continuum over the lines 0.55 0.65 0.81 0.97 1.06 1.06 1.06 1.68 1.68 1.27 1.27 1.93 0.87 0.90 0.95 1.01 1.04 1.04 1.04 1.23 1.23 1.06 1.06 1.37 1.03 0.96 0.84 0.72 0.70 0.70 0.70 0.59 0.59 0.95 0.95 0.40 1.04 0.96 0.84 0.71 0.70 0.70 0.70 0.58 0.58 0.94 0.94 0.38 Continuum Colors 4.4 4.4 4.4 4.4 9.7 9.7 9.7 10.5 10.5 4.5 4.5 0.0 0.44 0.56 0.76 0.96 1.05 1.05 1.05 1.79 1.79 1.40 1.40 2.11 Hγ 4.9 4.9 4.9 4.9 7.7 7.7 7.7 9.9 9.9 3.5 3.5 0.0 0.58 0.67 0.82 0.97 1.06 1.06 1.06 1.65 1.65 1.25 1.25 1.90 Hβ 3.5 3.5 3.5 3.5 7.5 7.5 7.5 7.9 7.9 3.9 3.9 0.0 0.79 0.83 0.91 1.00 1.06 1.06 1.06 1.38 1.38 1.08 1.08 1.56 # 3290 3600 4020 4510 6630 7520 8700 1 2 3 4 5 6 7 8 9 10 11 12 0.17 0.27 0.42 0.57 0.71 0.71 0.71 1.10 1.10 2.10 2.10 3.78 0.27 0.37 0.52 0.67 0.65 0.65 0.65 0.94 0.94 1.52 1.52 2.56 0.39 0.52 0.74 0.95 1.04 1.04 1.04 1.84 1.84 1.45 1.45 2.20 0.71 0.77 0.88 0.99 1.08 1.08 1.08 1.52 1.52 1.11 1.11 1.73 1.01 0.99 0.96 0.92 0.92 0.92 0.92 0.82 0.82 0.95 0.95 0.74 1.01 0.96 0.89 0.81 0.81 0.81 0.81 0.68 0.68 0.99 0.99 0.56 1.04 0.96 0.84 0.71 0.70 0.70 0.70 0.58 0.58 0.94 0.94 0.37 Table 2. Observables in the base, which make the matrices W ⋆ 5870 A. ij . All continuum fluxes are normalized to the continuum at ij , F ⋆ ij and C ⋆ galaxy spectra, they have to be postulated in the synthesis of test galaxies performed below. An alternative and poten- tially more attractive approach to deal with the errors is to marginalize eq. (8) over σ. This yields more conservative es- timates for the parameters, insofar as they are independent of the actual observational errors. In this paper we follow the traditional approach of treating σ as a relevant constraint in the synthesis process. In order to insure a realistic correspondence between the values of the errors and the actual quality of the data 6 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao we have adopted a prescription for the σ's based on our experience in dealing with the measurement of W 's and C's in galaxy spectra (Cid Fernandes, Storchi-Bergmann & Schmitt 1998). Two sources of uncertainty affect the mea- surements of W 's: (1) the noise within the line window, and (2) the uncertainty in the positioning of the pseudo- continuum fluxes Ck at the pivot wavelengths, which affect Wj because they define the continuum over the line. The effect of the first of these sources upon σ(Wj) is straight- forwardly obtained through standard propagation of errors by specifying the noise level at λj , the spectral resolution δλ, and the size ∆j of the pre-defined window over which Wj is measured. In Bica & Alloin (1986a) system the pseudo continuum is defined interactively by visual inspection of the spectrum around the pivot λ's, which makes the es- timation of σ(Ck) non straight-forward. In Cid Fernandes et al. (1998) we found these errors could be roughly scaled from the signal-to-noise ratio measured in nearby 'line-free' regions, such that the S/N in the pseudo continuum is (S/N )PC ∼ ρPC(S/N )λ, with ρPC ∼ 2–3. This leads to the following expression for the σ(Wj)'s: σ2(Wj) = ∆j δλ(S/N )−2 λ + (∆j − Wj)2ρ−2 PC(S/N )−2 λ (9) where the two terms correspond to the two sources of un- certainty discussed above. For simplicity, we further assume that the noise in the observed spectrum is constant for all λ's, which results in (S/N )λ = Cλ(S/N )5870. The errors in the Ck's thus become also constant: σ(Ck) = ρ−1 PC(S/N )−1 5870 (10) Besides yielding realistic values for the σ's, this recipe has the advantage of quantifying the quality of the data in terms of a single quantity: The signal-to-noise ratio at the normalization wavelength 5870 A. Of secondary importance are the spectral resolution and ρPC, which in the simulations below are kept fixed at 5 A and 3 respectively. In order to provide a quantitative notion of how a value of S/N translates into σ(Wj)'s, we have computed these quantities for the observables corresponding to spec- tral groups S1 and S7 of B88, typical of the nuclei of early (red) and late (blue) type spiral galaxies respectively. For S1 our error recipe yields σ(WCaIIK) = 36A/(S/N )5870 and σ(WMgI) = 19A/(S/N )5870, so errors of ∼ 1 A or less are achieved with (S/N )5870 ≥ 30. For S7 these lines have σ(W ) ∼ 15A/(S/N )5870. 3.3 Exploration of the parameter space: Sampling methods The exploration of the (x, AV ) space can be approached from two alternative perspectives: (1) A minimization prob- lem, employing methods to search for the global χ2 mini- mum (maximum likelihood). Minimization techniques were explored in S91 and Pelat (1997) for the synthesis of W 's only. (2) A sampling problem, where one seeks to map out the posterior probability distribution (eq. 7). The methods discussed in this paper are sampling meth- ods, which do not explicitly search for a minimum. In the limit of small errors, however, one expects the posterior maps to peak at the most likely model. For large errors, on the other hand, the best model sampled may be well off the true global minimum, but in this case the probability distri- bution is so broad that the very meaning of 'best model' is questionable. In such cases it is arguably more important to estimate plausible ranges for the parameters than to carry out a refined search for the most likely values. Our main motivation to explore the sampling approach is that most applications of EPS with Bica's base to date focused on the estimation of mean parameters, whose mean- ing cannot be directly ascertained without estimating their uncertainty and/or biases. A critical re-evaluation of this method is thus crucial to establish to which level of detail the results of this popular technique can be trusted. 3.3.1 Uniform sampling In order to compute the individual posterior probabilities for each parameter, we first note that eq. (8) for the xi posterior is simply the mean probability P (x, AV D, σ, H) over the space spanned by all possible values of xj6=i and AV for a fixed xi. The simplest method to compute such probabilities is to divide the (x, AV ) space into a uniform grid with ∆x steps for the population fractions and ∆AV for the reddening parameter. One then approximates eq. (8) by the finite sum P (xiD, σ, H) ≃ Ps e−E(xs ,AV,s)δ(xi − xi,s) (11) Ps e−E(xs ,AV,s) where s denotes a point (a 'state') in the 13-D grid and the δ(xi − xi,s) term retains only those grid points where population i contributes with fraction xi of the light. This discrete sweeping of the parameter space is easy to implement, and was in fact the recipe followed by most works with Bica's base to date (see section 1). A serious drawback of uniform sampling is its computational price. The number of grid cells is an extremely steep function of the resolution ∆x. A coarse ∆x = 10% sampling yields 3.5 × 105 x-points, increasing to ∼ 8 × 107 for ∆x = 5% and more than 4 × 1011 for a 2% resolution! To obtain the grid size one has to further multiply these numbers by the number of points in the AV grid, whose limits have to be pre-defined. Schmitt et al. (1999), for instance, performed a synthesis study with ∆x = 5% and ∆AV = 0.06 for AV between 0 and 1.5, which yields 2 × 109 grid points, perhaps the finest grid ever used in EPS with this 12 elements base. One obviously always aim for the best resolution possible, but a priori estimates of what resolution is necessary are not straight-forward. A 'good' resolution should produce prob- ability distributions that are smooth over scales of ∆x and ∆AV . The width of these distributions is of course a func- tion of the observational errors and of the set of observables being synthesized. 3.3.2 Metropolis sampling A further drawback of uniform sampling is that the algo- rithm spends most of the time in regions of negligible prob- ability, which contribute little to the integral in equation (8), or its numerical version (eq. 11). A more efficient sweep of the parameter space would be to use a sampling scheme which traces the full probability distribution (eq. 7). One Empirical Population Synthesis: Sampling methods and tests 7 such 'importance sampling' scheme is the Metropolis algo- rithm (Metropolis et al. 1953), which samples preferentially regions where the posterior is large. This is the parameter- space exploration method adopted in this work. Our implementation of the Metropolis sampler was as follows. Starting from an initial arbitrary point, at each it- eration s we pick one of the 13 variables at random and change it by an uniform deviate ranging from −ǫ to +ǫ, producing a new state s + 1. If the variable is one of the 12 xi's, the whole set is renormalized to unit sum. Moves to- wards unphysical values (xi < 0 or > 1, AV < 0) were trun- cated. Downhill moves (i.e., towards smaller χ2) are always accepted, whilst changes to less likely states are accepted with probability e−(Es+1−Es ), thus avoiding trapping onto local minima. This scheme, widely employed in statistical mechanics (Press et al. 1992; MacKay 2001 and references therein), produces a distribution which tends to the correct one as the number of samples Ns increase. There is, how- ever, no universal prescription to choose optimal values for ǫ or Ns, so some experimentation is needed. 3.4 Performance tests In order to test our Metropolis EPS code we performed a series of simulations for artificial galaxies generated out of the base. Two sets of test galaxies were used. The first is composed of the 26 galaxies used by S91 (see their Table 5) to test their minimization algorithm. As they did not syn- thesize colors, we have used a fixed value of AV = 2/3 to redden all their galaxies. The second set is composed of 100 test-galaxies whose population vectors and extinctions were generated randomly using a scheme which insures that in many cases a few components dominate the light. All sim- ulations presented in this paper were performed sampling Ns = 108 states, with the 'visitation parameter' ǫ set to 0.005 for both the xi's and AV . Experiments were also per- formed changing these quantities, and found to give nearly identical results. We note that less samples can be used if one is only interested in the mean (x, AV ), as this naturally converges much faster than its full probability distribution. Simulations were performed for severals values of (S/N )5870 and different sets of observables. Here we restrain the dis- cussion to the illustration of the sampling method. In Fig. 1 we plot the individual projected posterior probability distributions for each of the 12 xi's plus AV for two of S91's test galaxies composed of widely different stel- lar populations. The posteriors were computed synthesizing all 9 lines and 7 colors in the base ('Set A', as defined be- low) and using (S/N )5870 = 1000 to define the errors in the observables. This is clearly an unrealistic, highly idealized situation, but it serves as a test-bed for the method in the limit of perfect data. The simulations were started from the xi = 1/12 and AV = 0.5 point. The plot shows that the posterior probabilities are sharp functions peaking at the expected input value, marked by upper arrows, demonstrat- ing that the Metropolis sampler converges to the most likely region. One notes, however, that the strong components in the old age bin (such as x1, x2 and x3 in the 'old galaxy' example, drawn as solid lines) have broader individual pos- teriors. This effect is further discussed in the next section. In the opposite limit of large errors, σ → ∞ (the 'infi- nite temperature' regime, as it would be called in statistical mechanics) the likelihood loses its discriminating power and becomes ∼ constant over the whole parameter-space. Eq. (8) then tends to P (xi) ∝ (1 − xi)10 for a n⋆ = 12 base. In this regime the data does not constrain the model at all, and this expression for P (xi) simply reflects the volume of the x-space spanned by a given value of xi. Not surprisingly, this distribution peaks at xi = 0, and has a mean value xi = 1/12. We have verified that the Metropolis scheme re- covers this distribution satisfactorily as S/N → 0, which serves as a further test of the adequacy of this sampling method. The skewing of several of the individual posteriors plotted in Figs. 3 and 8 towards small xi is partly due to this effect, which drags the components to an intrinsically larger region of x-space. The evolution of the Metropolis walk through the 13-D (x, AV ) space is illustrated in Fig. 2. For clarity, only 6 com- ponents of x and just one state for each 104 steps are shown. Note that projected P (xi) posteriors such as those in Fig. 1 are essentially histograms of the xi states sampled in walks as that in Fig. 2, but weighted by their exp −χ(x)2/2 like- lihood. The input values of the components shown are indi- cated at the right edge of the plot for comparison. This par- ticular example corresponds to a simulation with S/N = 60 and with all 16 observables synthesized. One sees relatively large excursions of the parameters from their true values. Furthermore, anti-correlations between some of the compo- nents (x1 and x2, x8 and x9) are clearly visible. This 'mirror effect' becomes much more pronounced as S/N increases, whereas for low S/N the broadening of the likelihood func- tion washes out such correlations. This is a consequence of redundancies between several of the base components, as further discussed below. The top panels in Fig. 2 show the evolution of the χ2 and how two of the synthesized observ- ables, W (CaIIK) and C4020, oscillate very much within their respective ±1 sigma error ranges. One therefore anticipates an excellent fit to the data, and indeed the χ2 obtained for the mean (x, AV ) 'solution' in this example is just 0.8 (see also Fig. 7). Overall, we conclude that the Metropolis algorithm is an efficient sampler of the parameter space for EPS prob- lems. It provides a much more continuous mapping of x and AV than would be possible with a uniform grid. Further- more, it is substantially more efficient computationally. A uniform grid with 108 points would only sample each xi at ∼ 7% steps and 10 values for AV . By concentrating on the relevant regions of the parameter space, we achieve a much better resolution, as can be judged by the smoothness of the posterior distributions on scales of 1–2%. One possible variation of this scheme is to combine the uniform grid and Metropolis routines, starting a Markov chain from each point on a uniform grid. Experiments with this alternative scheme were performed, and found to pro- duce essentially the same results. This is to be expected, since we verified that the starting point has little effect upon the resulting posterior distributions. In fact, we performed a series of simulations using the input parameters as the starting point and obtained identical results. Though further improvements are possible, the simple Metropolis sampler already provides a substantial improvement over previous works with this base. 8 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao Figure 1. Individual posterior probability distributions for each of the components of the population vector x and the extinction, computed for an old population dominated test galaxy (solid lines) and a 'younger' galaxy (dotted lines). The top arrows indicate the input values of the corresponding parameter in each panel. The numbering of the x components follows the convention in Table 1. All 16 observables in Bica's base (9 absorption lines + 7 continuum points) were synthesized in this example and a S/N of 1000 was adopted to illustrate the convergence of the Metropolis sampler in the limit of (almost) perfect data. 4 TESTS OF THE BASE performed for (S/N )5870 = 10, 30, 60, 100, 300 and 1000, and three different sets of observables: Having developed a probabilistic formulation of EPS and a new sampling method, we now present a series of numer- ical experiments designed to evaluate the actual ability to recover the stellar population mix of galaxies using Bica's base. The tests were performed with the same series of fic- tional galaxies described in section 3.4. Simulations were • Set A: All 16 observables in the base. • Set B: Only the 9 absorption line equivalent widths. • Set C: The W 's of CaII K, CN, G-band and MgI+MgH plus the continuum fluxes at 3660, 4020, 4510 and 6630, all normalized to 5870 A. Empirical Population Synthesis: Sampling methods and tests 9 Figure 2. Illustration of the 'step by step' Metropolis walk through the parameter space. The six bottom panels illustrate the evolution of selected components of the population vector (in %), whose input xi values are indicated at the right edge of each plot. The values of the equivalent width of CaIIK and of the color C4020 are plotted for each step in the third and second panels from the top. The dotted horizontal lines in these plots mark the ±1σ range around the observed value. The top panel shows the χ2 for each sampled state. For clarity, only the first few million steps are show, and only 1 for every 10000 sampled states is actually plotted. All 16 observables in the base and S/N = 60 were used in this example. Set A is the ideal, as it uses all information in the base. Set B is used as a test of how much is actually gained by synthesizing colors as well as W 's, while Set C is composed of the observables used by Cid Fernandes et al. (1998) and Schmitt et al. (1999) to characterize the stellar content of ac- tive galaxies and their hosts through long-slit spectroscopy. These data will be used to perform a spatially resolved EPS study in a future communication. 4.1 Effects of the errors in the observables We first investigate the effects of the measurement uncer- tainties in the observables upon the results of the synthesis. Qualitatively, one expects the degradation of the data qual- ity to broaden the probability distributions. Furthermore, a shift and skewing of the P (xi)'s of intrinsically large compo- nents towards small xi's is also expected for low S/N due to 10 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao intrinsically larger number of states in this region of x (the 'dragging effect' discussed in §3.4). This is exactly what is observed in Fig. 3, where we overplot the individual pos- teriors for S/N = 1000, 100 and 10 for one particular test galaxy of S91, synthesized with Set A observables. As in Fig. 1, the true parameters are indicated by the top arrows. The broadening of posteriors in this case can be safely at- tributed to the errors alone, since, as demonstrated by Pelat (1998) most of the test galaxies in S91 (including the one in Fig. 3) have a unique algebraic solution in a W -only EPS problem, despite the fact that the number of degrees of free- dom (11 if we do not model the colors) falls short of the number of observables (9 W 's). Despite the large number of constraints in Set A, one sees that the individual posterior distributions are substan- tially broad even for a S/N = 100 spectrum. The average standard deviation of the xi's for this combination of S/N and observables is 3% over all simulations, but can reach more than 10% in individual components. Furthermore, the dominant components, 2, 6 and 9 in this example, are all badly affected by the errors. Fig. 3 shows that decreasing data quality progressively shifts P (x2) towards smaller val- ues, whereas the opposite happens with x1. This compensa- tion is also seen between components in the 1 Gyr (x5, x6 and x7), 100 Myr (x8 and x9), and 10 Myr (x10 and x11) age bins and is the same 'mirror' effect observed before in Fig. 2. A compelling visualization of these compensations is given in Fig. 4, where we plot the individual Metropolis states for the test galaxy employed in Fig. 3 in a xi × xj matrix. The S/N of 300 in this plot was chosen for clar- ity purposes; the same structure is present for other values, with the scatter around the strongest correlations increasing steadily as S/N decreases, to the point that at S/N = 10 the anti-correlations between components 1 and 2, or 8 and 9 become clouds similar to x1 × x5 in Fig. 4. All anticorre- lations identified in Fig. 3 (x1 × x2, x6 × x5, etc.) are clearly represented in this graphical version of the correlation ma- trix. These effects are rooted in the internal structure of the base. By definition, the base elements must be linearly in- dependent, and the S91 base complies with this algebraic condition. Yet, some of its elements are practically linearly dependent, in the sense that combinations of other elements can recover them to a high degree of accuracy. As the noise increases, the residuals between representing an element by its exact W 's and C's and a combination including other ele- ments with more extreme values for the observables become statistically insignificant, explaining the compensations de- tected above. To demonstrate this we synthesized the base elements themselves with our code. Results for Set A ob- servables and three different values of S/N are shown in Fig. 5. Each panel corresponds to one of these 12 extreme test galaxies (labeled B1. . . B12), with the mean synthetic values of each of the components plotted along the vertical axis. The empty, shaded and filled histograms correspond to S/N = 100, 30 and 10 respectively. Some of the components, most notably numbers 2, 3 and particularly 6, are synthe- sized with large contributions of others even for S/N = 100. For the x6 = 100% model (top left panel of Fig. 5), for in- stance, we find x6 = 60% for this high S/N , with 35 of the remaining 40% redistributed among components 5 and 7. Yet, the 16 observables are very well reproduced by this combination, with a total χ2 of just 3.8. This spreading of the light fractions becomes more pronounced for lower S/N 's, to the point that at S/N = 10, components like 8 and 9, or 10 and 11 cannot be distinguished anymore, and end up dividing their contributions in roughly equal shares, with residuals spread over other components. The spread is much smaller in the synthesis of components 1, 4, 5 and 7, which represent extreme metallicities within age groups. For intermediate Z systems (models B2, B3 and B6), however, these extreme components attract much of the percentage light contribution spilled over from neighboring populations of same age, so that their individual contributions in a real stellar population mix are as unreliable as the others. This experiment demonstrates that for practical pur- poses the base is still linearly dependent, at least in a statis- tical sense, despite the effort of S91 in reducing the highly redundant original base of B88. This 'noise-induced', or 'sta- tistical linear dependence', as we may call it, could indeed be inferred from the PCA analysis of S91, which showed that very little of the variance is associated with the last few eigenvectors. Within our probabilistic formulation, this 'statistical linear dependence' defines the structure of the likelihood function, which spreads first along directions in x-space producing similar observables, carrying the mean x to more densely populated regions. This effect explains the redistribution of the probability observed in Figs. 3, 4 and all other simulations. It is therefore extremely difficult to retrieve accurately all the input stellar population parameters even for excellent data. This is further illustrated in Fig. 6, where the input parameters are plotted against their mean values, as derived from the Metropolis runs. One hundred artificial galaxies are plotted in each panel, with open circles corresponding to Set A observables and S/N = 1000, and crosses indicating the results for Set A and S/N = 60. A systematic underestima- tion of strong components is seen for S/N = 60, a good spec- trum by any standard. This happens at the expense of an overestimation of the weaker components, which produces the large scatter seen in the bottom left corners of Fig. 6. The average uncertainties in the xi's are of order σ(xi) ∼ 3– 4% for S/N = 60. These are not enough to account for the large differences between input and output xi's depicted in Fig. 6, so the errors induce a true bias in x. Note also that, in accordance with the discussion above, this bias is smaller for models with strong contributions of components 1, 4, 5 and 12, because of their extreme locations in the space of observables. Only for unrealistically large S/N 's, which far exceed the quality of the data used to build the base in the first place, one is able to break the 'statistical linear dependence' of the base by distinguishing fine details in the observables. This should not be interpreted as a failure of the method, as what the code actually synthesizes are the observables! These are very precisely reproduced by the mean (x, AV ), as illustrated in Fig. 7 for the W 's of CaII K, CN, G-band and MgI and two continuum colors. The remaining observables, not show in the figure, are equally well reproduced. Further- more, whilst S/N > 300 is needed to recover accurately the model parameters from the mean solution, the agreement between the synthetic and measured observables is excellent for any S/N . Obviously, an even better agreement is ob- tained if instead of (x, AV ) the best model sampled during Empirical Population Synthesis: Sampling methods and tests 11 Figure 3. Effect of S/N upon the individual probability distributions of the synthesis parameters for Set A observables. The different lines correspond to S/N = 1000 (solid), 100 (dotted) and 10 (dashed). Notice how the probability of x components with a large contribution (like x2, x6 and x9 in this example) are progressively redistributed among other components as the noise increases. the Metropolis excursion is used to reconstruct the observ- ables, as illustrated in the fourth column of plots in Fig. 7 (see §4.5). In conclusion, this comparison of input and output ob- servables shows that the spreading of the probability and the confusion between components seen in Figs. 3 and 5 is not an artifact of the method, but a consequence of the in- ternal structure of the spectral base. As found in other EPS studies (O'Connell 1996 and references therein) synthesiz- ing the observables is one thing, but trusting the detailed star-formation history and chemical evolution implied by the synthesis is an altogether different story. On the pos- itive side, a notable fact about Fig. 5 is that, consulting Table 1, one realizes that the reshuffling of the strength of the components occurs preferentially among populations of same age, an effect also clearly seen in Fig. 4. Though some redistribution among base elements of different age and Z in the older 1 and 10 Gyr bins due to the age-Z degeneracy also occurs, this indicates that the age distribution may be well recovered by the synthesis (§4.3). 12 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao Figure 4. 'Correlation Matrix' for the test galaxy in Fig. 3. Each xi ×xj panel shows one every 10000 of the first 107 steps of a Metropolis run for (S/N )5870 = 300 and Set A observables. The values at the bottom of each row indicate the input value of the corresponding component (x1 = 0, x2 = 20%, . . . x11 = 10%, x12 = 10%). These are also indicated by the solid circles in each panel. Small and big tick marks are spaced by 2 and 10% respectively in all plots. Note the strong anti-correlations between adjacent components of same age. 4.2 Effects of different sets of observables The effects of which set of observables is used in the synthesis are illustrated in Fig. 8, where results for Sets A, B and C for the same test galaxy as in Fig. 3 and S/N = 1000 are compared. The figure shows that it is very advantageous to synthesize colors along with equivalent widths, despite the fact that one increases the dimensionality of the problem with the inclusion of AV . This can be seen by comparing the performances of Sets A and B. The information contained in the colors not only improves the estimation of x but also allows a very good determination of AV , which, unlike the population vector, is always well recovered by the synthesis. Fig. 8 also shows how important it is to provide as many observational constraints as possible to constrain the syn- thesis, as Set A recovers the parameters much better than Sets B and C. Decreasing the number of observables in the synthesis thus have the same overall effect as decreasing the S/N . The width of the posteriors for Set C are partially attributed to the algebraic degeneracy in this set (8 observ- Empirical Population Synthesis: Sampling methods and tests 13 Figure 5. Synthesis of the base elements. Each panel shows the mean synthetic population vector (vertical axis) for 12 test galaxies whose 16 observables were generated with xi = 100%, i = 1, . . . 12 (labeled B1. . . B12 respectively). Empty, shaded and filled histograms correspond to S/N = 100, 30 and 10 respectively. Ideally, all histograms should be concentrated in the component used to generate the observables. In practice, some bases elements, most notably 2, 3 and 6, are well synthesized by combinations of others even for large S/N . ables and 12 degrees of freedom), which implies the exis- tence of exact solutions in a sub-space of (x,AV ). That is not the case of Set A for the test galaxy in this example (Pelat 1998), but there, as for other sets, the 'statistical lin- ear dependence' of the base is very efficient in spreading the likelihood even for such an idealized S/N ratio. The excellent agreement between the 'observed' and synthesized observ- ables even for much worse S/N (Fig. 7) illustrates this point. Algebraic degeneracy is not as critical as the 'statistical de- generacy' induced by the combined effects of noise, limited data and the internal correlations in the base. Studies which synthesize only W 's sometimes use the resulting population vector to compute a predicted contin- uum shape and derive AV a posteriori through the compar- ison with the observed continuum (e.g., B88; Boisson et al. 2000). From our Set B simulations, we find that this pro- cedure yields values of AV 2–3 times less accurate than ob- tained with Set A. For S/N = 60, for instance, Set A yields 14 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao Figure 6. Input × output plots for all 13 individual parameters in the synthesis. Open circles map the results for Set A and S/N = 1000, while crosses correspond to Set A and S/N = 60. The output parameters are the mean values sampled from the global likelihood function. The dotted, diagonal lines in each panel mark the 'input = output' line. Error bars on the model parameters are not shown for clarity. Notice the systematic underestimation of large xi's and the corresponding overestimation of weak proportions for S/N = 60. an rms dispersion in AV of 0.06 mag around the true value, whereas the dispersion for set B is 0.13 mag. Since this is a relatively small difference, these experiments validate the a posteriori computation of the extinction. Colors are more useful to constrain the population vector than to estimate AV . 4.3 Age grouped results The results of the tests above reveal a striking difficulty to accurately determine all 12 base components of Bica's base in the presence of even very modest noise and/or when not all base observables are available for the synthesis. As in- dicated by the simulations, both the measurement errors and the use of reduced sets of observables act primarily in the sense of spreading a strong contribution in one compo- nent preferentially among base elements of same age. Group- Empirical Population Synthesis: Sampling methods and tests 15 Figure 7. Input × output observables for a series of 100 test-galaxies synthesized with the Metropolis-EPS sampler. The three left columns of plots correspond to the three indicated combinations of the set of observables and S/N . For these plots the x-axis quantities were computed with the mean (x, AV ) solution. The plots on the right column were made with Set C and S/N = 30 (as those in the third column), but the synthetic observables in this cases were computed with the best model found during the Metropolis runs. Only the equivalent widths of four absorption features and two colors exhibited, but results for the remaining observables are equally good. The vertical and horizontal scales are the same for each panel, and the input = output is indicated by a dotted line. Notice the excellent agreement, despite the fact that the input parameters are not accurately recovered by the synthesis process. ing the population vector in age bins should thus produce more robust results. This expectation is confirmed in Fig. 9, where, analogous to what was done in Fig. 6, we plot the input values of the xi's against the output xi's, but now for the five age binned groups, also for Set A simulations. The left panels correspond to the same values of S/N used in Fig. 6, so that one can appreciate the enormous improve- ment achieved by grouping equal age elements. The right panels show that good results are also obtained for S/N of 30 (circles), and that young populations (≤ 100 Myr) are reasonably well traced even for S/N as low as 10 (crosses). Though some deviations survive, specially for the smaller S/N 's and the older age bins, the age-binned synthesis re- sults are much more reliable than the results for the indi- 16 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao Figure 8. Analogous to Fig. 3, but exploring the effect of different sets of observables upon the synthesis parameters. Solid lines correspond to Set A (all observables), dotted lines to Set B (only 9 equivalent widths) and dashed lines to Set C (4 lines and 4 colors). vidual components, and, most importantly, do not require unrealistic S/N . This conclusion also holds for Sets B and C, as illus- trated in Fig. 10. Naturally, the restricted input data in these sets translates into a loss of information, and the agreement is not as good as for Set A. Note that Set C does a much better job in recovering populations of 100 Myr or less than Set B, a further example of how useful colors are in the syn- thesis. This happens because the color information (present in Set C but absent in B) impose stronger constraints on young populations than in older systems. The reliability of age-grouped proportions contrasts with the badly constrained and biased results for the in- dividual components of x, and is one of the main results of our study. This result puts in perspective all previous inter- pretations of the synthesis with this base. This fundamental limitation of the base was in fact previously known, but was never studied in detail. Schmitt et al. (1999), for instance, preferred to summarize the results of their synthesis study in terms of age grouped proportions for populations younger than 1 Gyr, in consonance with the results above. Still, in that work we kept a distinction between the different Z's in Empirical Population Synthesis: Sampling methods and tests 17 Figure 9. Age-binned light fractions in 100 input test galaxies versus the corresponding synthetic means. Left panels: Circles correspond to S/N = 60, and crosses to S/N = 1000, as in Fig. 6. Right panels: Circles correspond to S/N = 30, and crosses to S/N = 10. All results are for Set A observables. the 10 Gyr age-group. At this level of detail the population fractions are very uncertain. In fact, since Set C observables and S/N ∼ 30–60 were used in that study, Fig. 10 indicates that it would be safer to group their 1 and 10 Gyr pro- portions. Yet, the proportions for the younger (≤ 100 Myr) populations found by Schmitt et al. (1999), and which con- stituted the focus of that paper, are trustworthy (Fig. 10). 4.3.1 The age × metallicity degeneracy A close inspection of the S/N = 10 and 30 results in Fig. 9 reveals a compensation effect between the 1 and 10 Gyr age bins, albeit at a much smaller level than for the individual components (Fig. 6). This effect is much more pronounced in Sets B and C, as seen in Fig. 10. The confusion between the 1 and 10 Gyr populations is related to the age-Z degeneracy, which sets in precisely in this age range. This well known effect (O'Connell 1986, 1994; Worthey 1994) is also present in Bica's base and affects 18 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao Figure 10. Input × output age binned x fractions for Set B (left) and Set C (right) observables. Circles correspond to S/N = 100, and crosses to S/N = 30. the synthesis results. To visualize and quantify this effect we have computed the average log Z and log age for both the input and synthetic x for our test galaxies. Given that x is a light fraction at 5870 A, these averages ultimately represent flux-weighted ages and metallicities, which serve our current illustrative purposes, despite their ambiguous physical significance. The 'output - input' log age and log Z residuals are plotted against each other in Fig. 11. In these diagrams ∆ log age > 0 (overestimated age) implies redder colors and stronger metal lines, which tend to be counter-balanced by the bluer colors and weaker lines resulting from ∆ log Z < 0 (underestimated Z). The age-Z degeneracy is most obvious in the results for Sets B and C, for which the biases in age and Z persist even for S/N = 100. For Set A, on the other hand, the residuals are small: < ∼ 60, which shows that the age-Z degeneracy can be broken, at coarse the level of age and Z resolution offered by the base, with enough information and good spectra. Note that this remark applies to the global averages defined above, not the detailed component by component ages and Z's, which we concluded to be highly uncertain (Fig. 6). ∼ 0.2 dex for S/N > Empirical Population Synthesis: Sampling methods and tests 19 Figure 11. The age-metallicity degeneracy. The x-axis maps the difference between the output and input log age, with the corresponding difference in log Z plotted along the y-axis. Circles correspond to age and Z computed from the mean (x, AV ), whereas dots correspond to the best model sampled. All 126 test galaxies are plotted. The different panels correspond to combinations of the data quality (S/N , labeled in the bottom-left corners) and the set of observables (top-right) used in the synthesis. 4.4 Metallicity grouped proportions The results for Z binned proportions, plotted in Fig. 12, are not nearly as good as those for the age binned groups. With the exception of the Z = 10−2Z⊙ bin, which is actu- ally represented by just one element (Table 1), the scatter in the input × output Z-binned proportions is so large for all other Z's that one is forced to conclude that no accu- rate description of the chemical history of galaxies can be afforded with this spectral base. Given this limited Z 'res- olution power', one might consider removing intermediate components such as 2, 3 and 6. However, as we concluded that only the age distribution can be assessed with the base, there is no obvious advantage in this further reduction. This situation can be improved by providing a priori constraints on the occupancy of the age-Z plane, like those imposed by B88 based on chemical evolution arguments, since these ef- fectively reduce the dimensionality of the base to typically 8 components, thus alleviating the confusion between compo- nents and producing better focused results. The validity of 20 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao mass' of acceptable solutions. The severe biases identified in the mean synthetic population vector, however, raise the question of whether alternative approaches should be pur- sued instead. By definition, minimization oriented techniques would do a better job at recovering the input parameters. Indeed, we verified that the best model population vector sampled in the Markov chain is much closer to xtrue than its expected value, an agreement which can only be improved given that the Metropolis sampler is not an optimal minimization pro- cedure. Though the simulations confirmed our expectation that for small errors the mean and best solutions should both converge to the true parameters, it was surprising to realize how easily x moves away from xtrue, and consequently also from xbest. Very small uncertainties in the observables, at the levels of S/N > 300, are enough to set powerful com- pensation effects into operation (e.g., Figs. 3 and 4), shifting the components of x away from their true values due to the asymmetric redistribution of the probability in x-space. Minimization procedures can in principle overcome this bias, as illustrated by the success of the S91 and Pelat (1997) tests as well as by our (not optimized) results for xbest. A further bias in mean solutions is that they always produce xi > 0. Real galaxies, even if synthesizable by the base, are likely to have observables outside synthetic domain due to measurement errors. As shown by Pelat (1997, 1998), the best model in this case lies at the surface of the synthetic domain, where very often several of the x-components are 0. In these aspects, minimization algorithms are more suit- able than the more traditional sampling approach. However, the high susceptibility of the synthesis to the data quality indicates that minor perturbations in the input observables, within their errors, may seriously affect the estimation of the best model parameters. The stability of the best solu- tion is thus likely to be more fragile than that of the mean solution, which takes into account the effects of the errors in the observables in the trade-off between an optimal and a statistically acceptable fit of the data. We have in fact detected this effect in a series of simulations, but a more careful study, with specialized minimization techniques is required to quantify it properly. It would be particularly in- teresting to carry out this test with the method developed by Pelat (1997) applied to the base used in this work. His Monte Carlo tests with a n⋆ = 10 stellar base and nW = 19 equivalent widths produced essentially unbiased population fractions even for S/N as low as 10. Extending his formalism to synthesize galaxy colors would also be desirable. In any case, because of the nature of the EPS problem and the structure of Bica's base, relatively large discrepan- cies in the parameters do not necessarily translate into large differences in the synthesized quantities. In this respect, the difference between mean and best models is largely aca- demic, as the mean solution already does an excellent job in fitting the observables (Fig. 7). 5 SUMMARY In this paper we have revisited the method of Empirical Population Synthesis with the aims of improving it both at the formal and computational levels, and exploring its Figure 12. Metallicity binned light fractions in 100 input test galaxies versus the corresponding synthetic means. Circles corre- spond to S/N = 100, and crosses to S/N = 30, both for Set A observables. The 'resolution power' of the base is much better for ages (Fig. 9) than for metallicities. this prior input external to the synthesis process, nonethe- less, has to be evaluated in an object by object basis. In this study we followed S91 in not imposing any such extra information, as this approach encompasses a larger class of evolutionary scenarios and thus contemplates more possible applications. 4.5 Best × Mean Parameters Throughout these experiments, we have consistently used the mean parameters as an estimate of the result of the synthesis. This convention was deliberately chosen to follow the more widely employed version of EPS with this base and thus to allow a reassessment of previous results. Fur- thermore, the mean is a convenient and reasonable way to summarize the results insofar as it represents a 'center of Empirical Population Synthesis: Sampling methods and tests 21 efficacy as a tool to probe the population mixture of galaxies. Our results can be divided into three parts. (1) A simple probabilistic formulation of the problem was presented, which puts this method, traditionally em- ployed in a less formal way, onto a mathematical footing. It was shown how former applications of the method fit into the probabilistic formulation, thus providing some a poste- riori justification for previous results. (2) An importance sampling scheme, based on the Metropolis algorithm, was developed and tested. It provides a more efficient and smooth mapping of the probability dis- tribution of the parameters than is possible with the com- monly used uniform grid sampling. (3) In the third part of the paper, we applied the for- mulation and sampling method to a series of test galax- ies constructed out the n⋆ = 12 star-clusters base of S91. The tests explored the ability to reconstruct the model pa- rameters, which carry a record of the star formation and chemical evolution in galaxies, in the presence of (i) obser- vational errors and (ii) limited data. This study centered on the comparison of input parameters with the mean synthetic solution (x, AV ), as this is the most common form of EPS in the literature with this base, and a systematic evaluation of the consistency of this method has not been carried out before. The main results of this study can be summarized as follows: (a) The allowance for errors in the observables sparkle linear dependences within the base elements. This induces systematic biases in x, redistributing the probability in com- ponents with a large fractional contribution to the integrated spectrum among similar components. This happens prefer- entially for components of same age. Though this was a pre- dictable result, it was surprising to realize how powerful this effect is. For any realistic S/N , the bias is such that in most cases one cannot trust the estimates of all 12 individual com- ponents of x to any useful level of confidence. (b) Reducing the number of observational constraints, due to incomplete coverage of the 3300–9000 A spectral in- terval spanned by the base, has the same overall effect as reducing the data quality, that is, to broaden, shift and skew the probability distribution of the xi's. We find that synthe- sizing colors as well as equivalent widths yields substantially better results than synthesizing W 's only. The need to ac- count for the extinction as an extra parameter is largely compensated by the better focused x obtained by consider- ing the color information, particularly for populations of 100 Myr and younger. Unlike for the individual population frac- tions, we do not detect systematic biases in the estimates of AV . (c) Despite the difficulty to retrieve accurately the de- tailed population vector, the observables are very precisely reproduced by the mean solution. The mapping between the parameters and observables spaces is thus highly 'degener- ate'. The nature of this 'non-uniqueness' is not algebraic, as it happens also when the number of free parameters exceeds the numbers of observables and in undetermined cases where the solution is known to be unique. Instead, it is related to the possibility of mimicking to a statistically indistin- guishable level the effects of certain base elements by linear combinations of other elements. This happens even for tiny observational errors, inducing a 'spill over' of the likelihood among base components, carrying the mean solution along. (d) The structure of the base reflects the fact that evo- lution is the main source of variance in stellar populations. Accordingly, we found that grouping the x components in their five age bins produces much more reliable results, and we recommend this procedure when evaluating the detailed population vector obtained in previously published results with this base. Some confusion between the 1 and 10 Gyr populations also takes place due to the age-Z degeneracy, but this can be minimized with an adequate spectral cover- age (measuring all observables in the base) and S/N >∼ 30. Grouping the results for the four different Z's in the base is also recommended, but caution must be exerted when in- terpreting the results since these are much less precise than the age-binned proportions. The difficulties to retrieve accurately the true stellar population parameters stem from two sources: (i) The con- vention to represent the results of the synthesis by the mean solution, and (ii) limitations imposed noise-induced linear dependences in the base. Minimization techniques may be instrumental in overcoming the biases in the mean popula- tion vector, which ultimately limit the resolution of the 12-D base to its 5 age bins. Within our probabilistic formulation, other priors can be investigated. Maximum Entropy, for ex- ample, is a prior that has been applied successfully to inverse problems formally similar to population synthesis, like im- age restoration. However, more than improvements on the method, which are possible and desirable, further work on the base itself will certainly be needed. Indeed, the limited age and Z 'resolution power' of the base employed here lies at the heart of all difficulties identified in this paper. In- cluding other spectral indices, modeling the full spectrum and incorporating other relevant information such as mass to light ratios in the synthesis process are possibles ways to progress in this direction. ACKNOWLEDGEMENTS We thank Catherine Boisson, Monique Joly, Jihane Moul- taka and Didier Pelat for stimulating discussions on the virtues and drawbacks of EPS techniques. RCF thanks the hospitality of Johns Hopkins University, where this work was finalized. Support for this work was provided by the National Science Foundation through grant # GF-1001- 99 from the Association of Universities for Research in Astronomy, Inc., under NSF cooperative agreement AST- 9613615. JRSL acknowledges an MSc fellowship awarded by CAPES. Partial support from CNPq, FAPESP, PRONEX and FUNPESQUISA-UFSC is also acknowledged. REFERENCES Bica E., 1988, A&A, 195, 76 Bica E., Alloin D., 1986a, A&A, 162, 21 Bica E., Alloin D., 1986b, A&AS, 66, 171 Bica E., Alloin D., 1987, A&A, 186, 49 Bica E., Alloin D., Schmidt A. A., 1990, A&A, 228, 23 Bica E., Alloin D., Schmitt H. R.,1994, A&A, 283, 805 Bonatto C., Bica E., Alloin D., 1995, A&AS, 112, 71 22 R. Cid Fernandes, L. Sodr´e., H. Schmitt & J. Leao Boisson C., Joly M., Moultaka J., Pelat D., Serote Roos M., 2000, A&A, 357, 850 Bonatto C., Bica E., Pastoriza M. G., Alloin D., 1996, A&AS, 118, 89 Bonatto C., Pastoriza M. G., Alloin D., Bica E., 1998, A&A, 334, 439 Bonatto C., Bica E., Pastoriza M. G., Alloin D., 1999, A&A, 343, 100 Bonatto C., Bica E., Pastoriza M. G., Alloin D, 2000, A&A, 355, 99 Bruzual G., Charlot S., 1993, ApJ, 405, 538 Charlot S., Bruzual G., 1991, ApJ, 367, 126 Cardelli J. A., Clayton G. C., Mathis, J. S. 1989, ApJ. 345, 245 Cid Fernandes R., Storchi-Bergmann T., Schmitt H. R., 1998, MNRAS, 297,579 De Mello D. F., Keel W. C., Sulentic J. W., Rampazzo R., Bica E., White R. E. III, 1995, A&A, 297, 331 Faber S. M., 1972, A&A, 20, 361F Fioc M., Rocca-Volmerange B., 1997, A&A, 326, 950 Guiderdoni B., Rocca-Volmerange B., 1987, A&A, 186, 1 Jablonka P., Alloin D., 1995, A&A, 298, 361 Jablonka P., Alloin D., Bica E., 1990, A&A, 235, 22 Joly M., 1974, A&A, 33, 177 Kong X., Cheng F. Z., 1999, A&A, 351, 477 Larson R. B., Tinsley B. M., 1974, ApJ, 192, 293 Larson R. B., Tinsley B. M., 1978, ApJ, 219, 46 Leao J. R. S., 2001, MSc Thesis, Universidade Federal de Santa Catarina Leitherer C. et al. , 1996, PASP, 108, 996L Leitherer C., Schaerer D., Goldader, J. D., Delgado, R. M., Gonz´alez R., Carmelle K., Denis F., de Mello D. F.; Devost D., Heckman T. M., 1999, ApJS, 123, 3 MacKay D., 2001, Information Theory, Pattern Recognition and Neural Networks, http://wol.ra.phy.cam.ac.uk/mackay/ Metropolis N., Rosenbluth A., Teller A., Teller E., 1953, J. of Chem. Phys., 21, 1087. Moultaka J., Pelat D., 2000, MNRAS, 314, 409 O'Connell R. W.,1976, ApJ, 206, 370 O'Connell R. W., 1986, in Stellar Populations, ed. C. Norman, A. Renzini, M. Tosi (Cambridge: Cambridge Univ. Press), p.213 O'Connell R. W., 1994, in Nuclei of Normal Galaxies: Lessons from the Galactic Center, ed. R. Genzel (Dordrecht: Kluwer), O'Connell R. W., 1996, in From Stars to Galaxies: The impact of stellar physics on galaxy evolution, eds. C. Leitherer, U. Fritze-von-Alvensleben, J. Huchra (ASP: San Francisco), Pelat D., 1997, MNRAS, 284, 365 Pelat D., 1998, MNRAS, 299, 877 Pickles A. J., 1985, ApJ, 296, 340 Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B. P., 1992, Numerical Recipes, 2nd. edition (Cambridge: Cam- bridge University Press) Raimann D., Bica E., Storchi-Bergmann T., Melnick J., Schmitt H. R., 2000, MNRAS, 314, 295 Serote Roos M., Boisson C., Joly M., Ward M., 1998, MNRAS, 301, 1 Serote Roos M., Boisson C., Joly M. 1996, A&A, 117, 93 Schmidt A. A., Copetti M. V. F., Alloin D., Jablonka P., 1991, MNRAS, 249, 766 Schmidt A. A., Bica E., Alloin D., 1990, MNRAS, 243, 620 Schmidt A. A., Bica E., Dottori H. A., 1989, MNRAS, 238, 925 Schmitt H. R., Bica E., Pastoriza, M. G., 1996, MNRAS, 278, 965 Schmitt H. R., Storchi-Bergmann T., Cid Fernandes R., 1999, MNRAS, 303, 173 Silva D., Cornell M., 1992, ApJS, 81, 865 Spinrad H., Taylor B. J., 1971, ApJS, 22, 445 Tinsley B. M., 1972, A&A, 20, 383 Worthey G., 1994, ApJS, 95, 107 This figure "fig4.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0102061v1
astro-ph/0308538
1
0308
2003-08-29T15:45:27
Astrophysical imaging with the Darwin IR interferometer
[ "astro-ph" ]
The proposed infrared space interferometry mission Darwin has two main aims: (i) to detect and characterize exo-planets similar to the Earth, and (ii) to carry out astrophysical imaging in the wavelength range 6 - 20 micron at a sensitivity similar to JWST, but at an angular resolution up to 100 times higher. In this contribution we will first briefly discuss the imaging performance of the Darwin mission. We will then discuss how Darwin will contribute in a very significant way to our understanding of the formation and evolution of planets, stars, galaxies, and super-massive black-holes located at the centers of galaxies.
astro-ph
astro-ph
1 ASTROPHYSICAL IMAGING WITH THE DARWIN IR INTERFEROMETER H. J. A. Rottgering 1, L. d'Arcio1,2, C. Eiroa1,3,4, I. Labb´e1, and G. Rudnick5 1Sterrewacht Leiden, P.O. Box 9513, 2300 RA Leiden, The Netherlands 2Space Research Organisation Netherlands, The Netherlands 3Universidad Aut´onoma de Madrid, Spain 4ESTEC Noordwijk, The Netherlands 5MPA, Garching, Germany ABSTRACT The proposed infrared space interferometry mission Darwin has two main aims: (i) to detect and char- acterize exo-planets similar to the Earth, and (ii) to carry out astrophysical imaging in the wavelength range 6 - 20 micron at a sensitivity similar to JWST, but at an angular resolution up to 100 times higher. In this contribution we will first briefly discuss the imaging performance of the Darwin mission. We will then discuss how Darwin will contribute in a very sig- nificant way to our understanding of the formation and evolution of planets, stars, galaxies, and super- massive black-holes located at the centers of galaxies. Key words: frared: general; stars: galaxies: evolution. instrumentation: interferometers; in- formation; galaxies: active; 1. INTRODUCTION It will be very dim: The main focus of this conference is to discuss the wonderful recent results on planets orbiting other stars and the prospects of the Darwin and TPF mis- sions for detecting and characterizing planets with properties similar to our beloved Earth. For a long list of reasons, detecting and studying exo-earth's is technically very challenging. Let us consider an exo- earth in the habitable zone of a G-type star at a distance of 10 pc. in the IR, it will have a flux density of 0.34 µJy (Beichman et al. 1999), and in the optical magnitude of V = 30. Furthermore, it will be located at a close angular separation of 100 mas from the parent star. To be able to detect and study such earth-like planets very sensitive observations at high angular resolution are clearly needed. An instrument that is capable of de- livering such exquisite measurements will be able to also do very important observations for general astro- physics. In this contribution we will concentrate on the promise of Darwin for such studies. In the wave- length region around 10 micron, Darwin will have a similar sensitivity to JWST, but with an angular resolution a factor of 10 − 100 larger. The Darwin mission was an integral part of ESA's Horizon 2000+ programme and was designed from the beginning with these two goals in mind. The proposal by L´eger et al. (1996) to build Darwin originated from a time before the first exo-planet was discovered. With more than 100 planets now discovered, the landscape of this field has changed dramatically, making it even more clear that a ma- chine that can study exo-earth's is needed. Also for "normal" astrophysics the scene has changed signif- icantly, mainly due to the advent of 10-m class opti- cal/near IR telescopes and X-ray and infrared satel- lites. There are a number of very good reasons to have a Darwin mission with astrophysical observations as one of its key drivers. First, it is a logical next step after a long series of IR missions (IRAS, ISO, SIRTF and JWST). The angular resolution that will be needed as a follow-up of sources studied by SIRTF and JWST can not naturally be obained with a sin- gle dish telescope, but will require interferometry. Second, an important argument is that of "cost effec- tiveness": the Darwin mission will be very expensive, and if it can do "two missions at (more or less) the price of one", this will be very valuable. Third, there is the argument of redundancy. The whole mission will be very challenging, and having more than one goal will help reducing the scientific and technical risk of the entire mission. In this presentation, we will first discuss the perfor- mance of the Darwin imaging mission. Subsequently we will briefly describe the research into three ar- eas of astrophysics (star formation, galaxy formation and evolution, and active galactic nuclei) that Dar- win will contribute to in a major way. 2. IMAGING CONSIDERATIONS 2.1. Sensitivity The proposed configuration for Darwin consists of 6 telescopes each with a diameter of 1.5 m and a central beam-combiner. The relevant systems at the tele- scopes and beam-combiner will be passively cooled to 40 K, so that the sensitivity will be limited by shot-noise from emission by the zodiacal background (Fridlund et al. 2001, Nakajima and Matshura 2001). This system should be a able to detect a point source of 2.5 µJy with a signal to noise of 5 in one hour of integration time. For sources that are more com- plex, assessing the integration time depends on the details of the morphologies. A simple rule of thumb that is used in the radio astronomical community, is that an object can be imaged in a certain inte- gration time provided that the total signal within the primary field of view is larger by about a fac- tor 50 − 100 than the expected off-axis rms noise in the final map. For Darwin the point-spread function (PSF) for a single telescope is 1.4 arc-sec at 10 mi- cron. This is the maximum field of view (FOV) if the beam combination is done in the pupil-plane. The image requirements for Darwin therefore are that a source can be "mapped" with an integration of one hour, provided that the total flux within ∼ 1 arc-sec is more than 25 − 50µJy. 2.2. Resolution and filling of the UV plane The maximum baselines that are foreseen are about 500 meter, which translates into a maximum resolu- tion of 4 mas at 10 micron. This is a factor of more than 100 higher than the nominal resolution at 10 micron of 350 mas of the 6.4 m mirror of JWST. Let us consider an experiment with baselines of up to 150 meters. With the resulting resolution and the FOV as determined by the PSF's of the individual telescopes, a final map will contain up to 100 × 100 individual pixels. This number of pixels is directly set by the ratio of length of the maximum baseline to the size of the FOV. For mapping very complex ob- jects it is essential that the number of visibilities that are measured in the UV plane is significantly greater than the number of parameters that is needed to de- scribe the image. Furthermore, these visibility mea- surements should be well distributed over the UV plane. This means that in the case that our complex object occupies a significant fraction of the 1002 pix- els, also on the order of 1002 visibility measurements will be needed. Per Darwin configuration, a maximum of 15 indepen- dent visibility points can be obtained. This immedi- ately indicates that a 1002 map would need visibil- ity measurements from 666 different configurations. This is not such a severe requirement if we take a closer look at the dynamics of the configuration. For a more detailed discussion, see d'Arcio et al. 2003. A basic approach to filling up the UV plane might 2 Figure 1. The coverage of the UV plane for the sin- gle expansion of the configuration to maximum base- lines of 500 meter coupled with a 60 degree rotation. From d'Arcio et al. 2003. be a single expansion coupled with a 60 degree ro- tation For the resulting UV-plane coverage, see Fig. 1. With the planned thrusters delivering 1 mN, the fastest reconfiguration cycle takes about 16 hours. The average baseline rate will be about 1.5 cm s−1. The maximum integration time for a single UV point is then set by the requirement that each UV point should not be significantly blurred by the movement of the array. For a telescope of 1.5 meter a move- ment of 15 cm in the UV plane is certainly tolerable, indicating maximum individual integration times of about 10 seconds. This would yield 15 * 16 hours * 3600 sec / 10 sec = 86400 different UV samples, far more than our simple required minimum of 666 samples needed for complex maps. Interestingly, this leads to the conclusion that the free flying telescopes allow for much "easier" filling of the UV plane than is possible with fixed ground based telescopes. Furthermore, the number of UV points needed is modest compared to conventional radio synthesis telescopes. For example, for the VLA the ratio of maximum baseline size over the size of the individual apertures is 27 km/25 m = 1000, a factor of 10 more than Darwin. For very large base- line interferometry with radio telescopes scattered all over the globe, the ratio is many orders of magnitude worse. 2.3. Co-phasing While the telescopes are moving it is essential that the system is able to integrate to obtain the desired visibility measurements. Note that since the aim is to perform imaging, both the amplitude and the phase of the visibilities need to be measured. To be able to do this, the array needs to remain co-phased while moving. In general, the science targets are too faint to give enough signal for co-phasing. Instead the light from an off-axis bright reference star is used. A system based on such a principle is currently being implemented at the VLTI under the name PRIMA, which stands for Phase-Referenced Imaging and Mi- croarcsecond Astrometry. For a detailed description of this instrument we refer to the contribution of Quirrenbach to these proceedings. d'Arcio et al. 2003 discussed the various options for implementing off-axis referencing. An important de- sign choice is related to how to multiplex the science and reference beam. Sending those two beams along two different path is complicated: it requires rela- tively large optics and the calibration of the differen- tial path lengths is arduous. Sending the two beams along a common path is therefore clearly preferred. Multiplexing the beams can in principle be done us- ing polarization or splitting up into time sequences. However, multiplexing in wavelengths seems simplest and has therefore been selected by the Alcatel study (Fridlund et al. 2000). Within such a scheme the K-band is used for fringe-tracking. At the expected limiting magnitude for fringe tracking of mk=17, al- most 100 % of the sky should be observable, pro- vided that the angular distance of the science target and the reference star is allowed to be maximally 1.5 arcminute. An issue is cross talk between the sci- ence and reference channels, especially at the shorter science-wavelengths. d'Arcio et al. (2003) showed that levels of a few per cent cross-talk can be toler- able and this is probably achievable with a proper design. The implementation of such a scheme studied by Al- catel involved modifying the nulling beam combiner. This however, strongly reduces the sensitivity, po- tentially by a factor of about 10 − 20 (d'Arcio et al. 2003). For imaging it would therefore clearly be advantageous to have a separate beamcombiner. 2.4. Extending the field of view In the case of pupil-plane combination, the field of view of Darwin will be small, of the order of 1 arc- sec. Mapping a larger field of view can always be done through a mosaic of different pointings. This is of course very expensive in time. A natural way of enlarging the field of view is to combine the beams in the image plane. This is somewhat complex since the pupils need to be remapped at the beam-combiner with the correct demagnification and relative dis- tances. Furthermore, the required optics are large in size. An interesting idea, which potentially simpli- fies the design is to have zoom optics installed at the telescopes, so that at the level of the beam-combiner, the pupils do not need to be reconfigured. In the im- plementation of such a design that d'Arcio and le Poole discuss in these proceedings, a 4k*4k detector would yield a variable field of view, ranging from 25 arcsec for baselines of 50 m down to 2.5 arcsec for baselines of 500 m. 3 3. SCIENCE WITH DARWIN 3.1. Diagnostics The wavelength region between 5 and 20 micron is very rich in diagnostics for a number of species, in- cluding molecules, ions, dust and late type stars (e.g. van Dishoeck 2000). As a nice illustration of this richness, we show in Fig. 2 the infrared spectrum of the famous starburst galaxy Circinus as observed by ISO from Moorwood et al. (1996). For molecules, rotational and vibrational lines can be observed and for ions, forbidden fine-structure lines. Combined, these can be used to map out the tem- perature, density, and metallicity, kinematics of gas at intermediate to cold temperatures (10 −10, 000 K) and moderate densities 10 − 106 cm−3. In this wave- length regime the continuum is often due to emis- sion from dust and Polycyclic Aromatic Hydrocar- bon (PAH) molecules. Studying the precise form of the continuum reveals information about the compo- sition and temperature structure of the various dust and PAH components. The spectral energy distribu- tion of a normal nearby galaxy peaks at a rest-frame wavelength of a few microns. For galaxies at high redshifts, this peak shifts to observed wavelengths of 5 − 10 microns. High resolution sensitive measure- ments at 5 − 10 microns are therefore very impor- tant for understanding the formation and evolution of high redshift galaxies. 4. STAR FORMATION Based on a wealth of observations a sketch of a sce- nario showing how (isolated low-mass) stars form seems to be in place (Shu et al. 1987). Also the suggested classification scheme for the various evolu- tionary stages seems commonly accepted. The for- mation of a star commences with the growth of con- densations in molecular clouds. When the density of a condensation reaches a critical value, a collapse sets in, on a time scale likely to be governed by the local sound speed. Subsequently, a protostar surrounded by an accretion disk forms, both deeply embedded within an envelope of infalling dust and gas. A strik- ing phenomenon during this phase is the occurrence of bipolar outflows. Gradually, the inflowing matter will fall more and more onto the disk rather than the star. Within the final stage, the disk might be fully dispersed, for example by an energetic outflow. Alternatively, the disk material might (partially) co- agulate and form one or more planets. Within the vast range of physical conditions that oc- cur during star formation, the high resolution imag- ing provided by Darwin will map the density, tem- 4 Figure 2. The infrared spectrum of the famous starburst galaxy Circinus as observed by ISO (from Moorwood et al. 1996) perature, metallicity and dynamics of young stellar objects. This is particular relevant for studies of the core inflow, the accretion disks, the location of the jet outflow and planet formation. For example, it is not well understood how the jet outflow originates from the accretion disk. How does this start? What keeps the jet stable? Mapping out the structure of the disc will be very important for constraining mod- els of planet formation. For a first impression of how many sources are avail- able for detailed study with Darwin, we have made a compilation of various ISOCAM surveys at 6.7 and 14.3 micron down to a completeness limit of ∼ 10−15 mJy of three starforming regions (Persi et al. 2000, Bontemps et al. 2001, Kaas, priv. comm.). In Table 1, we give for each of these regions their names, the estimated age of the systems, their distance and the total area surveyed. Furthermore, the total number of sources detected both at 6.6 and 14.3 micron is given and, if possible, the number of sources in each of the star formation classes I, II and III. This classi- fication is done on the basis of the 6.7 and 14.3 colour indices. For more details see Bontemps et al. 2001. To investigate whether these sources have the ap- propriate angular scale for observations with Dar- win, we use as a starting point the work of Monnier and Millan-Gabet (2002). They have compiled scale sizes at 1.6 and/or 2.2 µm for the circumstellar disks around young stellar objects (YSOs), mainly Herbig Ae/Be and T Tauri stars. They find that the inner radii of the disks range from 0.1 AU up to 10 AU and appear to be correlated with the luminosity of the star (see Fig. 3). If we now apply this rough cor- relation between luminosity and size to the Class II sources as presented in Tabel 1., we find that the ma- jority of these sources have estimated angular sizes between 10 and 500 mas. An alternative way to es- Table 1. Results from a compilation of 6.7 and 14.3 micron ISOCAM surveys. Region Age (Myr) Distance (pc) Observed Size (deg. sq.) N(6.6 + 14.3) N(Class I) N(Class II) N(Class III) Chameleon Ophiucus Serpens 5 160 0.57 103 46 19 0.2-3 160 0.7 211 16 123 77 0.1 260 0.13 124 19 44 9 timate characteristic angular sizes for such objects is to use models of proto-planetary disks. A family of such models have been developed by D'Alessio et al. (2001). These models solve the radiative transfer through a viscous dusty disk heated by the central region. The resulting synthetic SEDs can then be compared to observed SEDs. Mer´ın et al. (these proceedings) used these models to predict angular size scales for proto-planetary disks and find angular size scales in the range of ten to a few hundred mas, consistent with the observationally based estimate for the sources presented in Table 1. From these two estimates it is clear that a proper population study requires the sensitivity and angular resolution of the Darwin mission. 5 Figure 3. The scale sizes at 1.6 and/or 2.2 µm as determined from interferometric measurements of circum- stellar disks around young stellar objects (YSOs), mainly Herbig Ae/Be and T Tauri stars versus the luminosity of the parent star (from Monnier and Millan-Gabet 2002). 5. ACTIVE GALAXIES Active galaxies are among the most spectacular ob- jects in the sky. Their compact nuclei can be so lumi- nous that they can outshine an entire galaxy. These objects show many interesting observational charac- teristics (e.g. radio lobes, jets, narrow and broad emission lines, polarized light, X-ray continuum and line emission, see reviews by Antonucci 1993 and Urry and Padovani 1995). The relative importance of these observational characteristics varies dramat- ically from object to object, and a complicated phe- nomenological scheme has emerged to classify objects into many different types of AGN, including Sey1, Sey2, QSOs, QSRs, BLACs, radio galaxies, etc. A major issue in this field is how to understand all these different characterics on the basis of the fundamen- tal physical and geometrical properties of the inner heart of the AGN. Orientation is an important factor in determining the appearance of an individual ob- ject. Furthermore, the characteristics of the central massive black hole (including total mass and spin) are likely to be directly related to the total output of the AGN and the presence of well defined jets. And finally, over the lifetime of the AGN activity its ap- pearance can change dramatically (for example radio sources can grow up to a few Mpc, starting at the pc scale). This evolution of AGNs is also evident in the fact that at a redshift of z = 2 bright AGNs are 1000 times more frequent than today. Although it is clear that this indicates a close link between the formation of galaxies and the AGN phenomenon the reason for the existence of this AGN epoch is not understood. More direct evidence for this link between AGNs and galaxy formation is provided by the relationship between galaxy and massive black hole mass, increasing the significance of AGNs for the overall understanding of galaxy formation. In the currently popular and attractive "unified" model of active galactic nuclei, all AGNs contain the following components: A central black hole fed by an accretion disk, which is surrounded by an opti- cally thick obscuring torus and broad and narrow line regions. The orientation with respect to the line of sight determines whether the object is observed to have broad emission lines, originating from within the hot central hole of the torus, or whether the torus blocks this region from view, leaving only the unob- scured narrow line region visible. Although this picture is capable of explaining a large number of the differences between the various classes of AGN (e.g. Antonucci 1993, Urry & Padovani 1995), it is still a vigorous debate whether other mechanisms contribute to, or even dominate over, the scenario in which orientation and a putative torus play such a major role. It has even been argued that in a subset of AGNs, the main power-source is not the black-hole but supernova explosions produced within a central starburst region. It is not clear, however, if all AGNs have starburst regions, whether all star- burst galaxies contain AGNs nor what the causal re- lation is between these two phenomena. An important issues in understanding AGN uni- fication are therefore whether dust tori exist and whether the physics/geometry of such tori can be constrained well enough for many of the differences between AGNs to be understood. 6 Figure 4. Greyscale representation of the morphol- ogy at 10 micron according to the model for NGC 1068 of Granato et al. (1997). For this particular model the visual extinction is AV = 150 mag and the orientations is theta = 60◦ (Courtesy Bjon Hei- jligers) Figure 5. The 10 micron luminosity and physical scale D of dusty tori as a function of redshift. The solid lines are for observed angular scales of 1, 0.1 and 0.01 arcsec. The dotted lines are for an observed 10 micron flux density of 5 and 50 µJy. The compu- tations have been done for H0 = 75 km s−1 Mpc−1 and Ω = 1. A number of radiative transfer models for obscur- ing tori models have been proposed (e.g. Krolik and Begelman 1988; Pier and Krolik 1992; Efstathiou and Rowin-Robinson 1995; Granato et al. 1997; Manske et al. 1998; Wolf and Henning 1999). Within these models the radiative transfer through a torus with a given geometry, optical depth and dust grain composition is calculated. The model subsequently gives the emerging IR spectrum and the morphology as a function of wavelengths. It is very clear that Darwin will be able to map nearby tori in exquisite detail. Recently the fa- mous Seyfert 2 galaxy NGC 1068 was observed at the VLT interferometer using the 10 micron MIDI instrument (see ESO press release 17/03 at http://www.eso.org/outreach/press-rel/pr-2003/pr-17-03.html). This first detection of fringes of an extragalactic object is an important step for optical interferom- etry. The plan is that during 2003 of the order of 10 − 20 visibility points will be measured. This will give important constraints on the general shape of the NGC 1068 torus. However, compared to VLTI, Darwin will be a dramatic step forward. The number of visibilities that will be obtained will be 3 orders of magnitude higher, with a fairly uniform filling of the UV-plane out to 500 meter. This will be very important to map in detail substructure, warps, and dynamics of nearby tori. An interesting question is to what distance tori of AGN can be mapped. To investigate this, we will use the torus models of Granato et al. 1997. In these models, the inner radii of these tori, rin, are set by the distance from the central source at which the dust grains sublimate due to the strong nuclear radiation. This radius is larger for more luminous AGN. For the models of Granato et al., rin ∼ 0.5L1/2 46 pc, where L46 is the luminosity of the central optical UV emitter in units of 1046 ergs s−1. As a scale size of the torus D we will use 300 rin. In Figure 5, the diameter D and the 10 micron luminosity (which are directly coupled) are given as a function of z for angular scales of 1, 0.1 and 0.01 arcsec. The dotted lines correspond to the 10 micron luminosities as a function of z for 10 micron flux densities of 5 and 50 µJy. A relatively weak AGN such as NGC 1068 has a 10 micron luminosity of the order of 1.7 × 1031 erg s−1 Hz−1 and its modelled torus size is 60 pc. Up to redshifts of z = 1 − 2 such weak AGN are brighter than the nominal sensitivity for a one hour imaging observation (see Section 2.) of 25 − 50µJy. Also the nominal resolution at 10 micron of 20 mas is very adequate for imaging the tori at these redshifts. Brighter AGN can basically be mapped up to redshift of z = 10 − 20 (if they exist). This shows that Darwin can not only study the physics of dusty tori in our local universe, but also at large redshifts. This will give the unique opportunity to investigate how the properties of tori change with redshift and when and how these tori and their asso- ciated massive black-holes are built up at an epoch when galaxies are forming. 6. GALAXY FORMATION AND EVOLUTION With the advent of 10-m telescopes, it is now possi- ble to define and study large samples of very distant galaxies. This work has been pioneered by Steidel and coworkers. Using the "Lyman break technique", they have defined a sample containing of order a 1000 galaxies between 2.5 < z < 5 (e.g. Steidel et al. 1999). Other techniques to obtain samples of very distant galaxies are to carry out deep imaging us- ing narrow band filters to find Lyα and Hα emitting galaxies (e.g. Venemans et al. 2002; Kurk 2003) or preselect on very red J − K colours (Franx et al. 2003). Also direct spectroscopic follow-up of SCUBA galaxies (e.g. Chapman et al. 2003), radio galaxies (e.g. de Breuck et al. 2001) and X-ray emitters (e.g. Rosati et al. 2002) can yield significant samples of z > 2 objects. This amazing observational progress of the last 10 years has been accompanied by a large body of work to build reliable and robust models of galaxy forma- tion. The goal is to model the evolution of galaxies with, as the main input, the physical conditions as they existed in the very early universe. For a thor- ough account we refer to a few of the excellent ar- ticles and reviews that have been written (see for example: Rees and Ostriker 1997; White and Rees 1978; White and Frenk 1991; Cole 1991; Lacey and Silk 1991; Kauffmann et al. 1993; Lacey et al. 1993; Cole et al. 1994; Kauffman et al. 1994; Somerville et al. 2001; Moustakas and Sommerville 2002; Bell et al. 2003) Virtually all of the current models assume that the dynamics of the large scale mass distribution in the early universe is driven by the gravity exerted by some form of dark matter. At early times the den- sity fluctuations within this medium are modelled by a described distribution whose functional form depends on the physics in the early universe, the nature of the dark matter and a suitable choice of cosmological parameters. The evolution of the dark matter distribution can be studied analytically or with the help of N-body simulations. A second step is to include the baryonic gas and to follow its hy- drodynamic evolution. Gas dynamics, shocks and radiative heating and cooling all need to be part of the simulation to obtain a realistic multi-phase medium. The outcome of this kind of simulation is that a significant fraction of the gas cools and settles at the centers of dark matter halos. It is from that gas that the stars that will make up future galax- ies will form. Often it is then simply assumed that the rate at which the gas at the center of these halos forms stars is proportional to the total amount of gas present and inversely proportional to the dynamical timescale within the dark matter halo. The rate with which these proto-galaxies forms stars is limited due to both supernovae and stellar winds which blow gas out of the centers of these halos. The merging of galaxies greatly enhances the combined star forma- tion rate of both galaxies, possibly to a rate at which a very large fraction of the gas is transformed into 7 stars within a few dynamical timescales. Finally, the combination of the inferred star formation rate, an assumed initial mass function and the spectral evolu- tion of individual stars will then give the evolution of the integrated spectra of an individual galaxy. With this kind of modelling gross properties of the general galaxy population can be calculated. These proper- ties include the luminosity function of galaxies, the redshift distribution, the relative numbers of ellipti- cals and spirals, faint galaxy counts, the history of star formation. Also more detailed properties like the size distribution, and the Tully-Fisher relation can be predicted with a fair accuracy. What role can Darwin in this area? Maybe it is important to realise that even 10 years from now the issue of how galaxies form will still be one of the most important topics within the field of astron- omy. Or to quote Rees (1998), who states that in 10 years from now, "we will probably still be unable to compute crucial things like the star formation ef- ficiency, feedback from supernovae. etc -- processes that current models for galactic evolution are forced to parametrise in a rather ad hoc way". An essential constraint on these galaxy formation models will be the spatial structure of very distant galaxies at 1 − 2 micron restframe, the location of the peak of the spectral energy distribution of nearby galaxies. For z ∼ 5 galaxies, this region is redshifted into the spectral window within which Darwin will be observing. A crucial question is whether there are sufficient number of distant galaxies on the sky with appropri- ate angular sizes for Darwin to observe. To investi- gate this we will use results from the Faint InfraRed Extragalactic Survey (FIRES, Franx et al. 2000), which is a very deep infrared survey centered on the Hubble Deep Field South using the ISAAC instru- ment mounted on the VLT (Moorwood 1997). With integration times of more than 33 hours for each of the infrared bands J, H and K, limiting AB magni- tudes of 26.0, 24.9, and 24.5 respectivily are reached (Labb´e et al. 2003). A major advantage of observ- ing this field is that multicolour HST photometry is available. Using the multi color data, we estimate the distance of the objects using the photometric redshift tech- nique. This technique uses a fit to the observed SED with a linear combination of galaxy templates. The procedure is described in detail in Rudnick et al. 2001. and 2003. An example of fits that are ob- tained is given in Fig. 6. The galaxies shown are a mixture of red J-K galaxies and Lyman Break galax- ies all at a photometric redshift of z ∼ 3. For further details, we refer to Labb´e et al. 2003. For these high redshifts, the reddest observed colour, K-band, samples only the rest-frame V -band. To sample the SED in the restframe K-band observa- tions at the Darwin wavelength range are essential. This is important so that a complete census of the stellar population in these distant galaxies can be ob- 8 Figure 6. Sample of template fits to photometric data for 8 objects in the HDF-S. The photometric data are from the HDF-S observations and from the FIRES project. The derived template fitting yielded photometric redshifts between 2.3 < z < 3.6. For further information see Labb´e et al. (2003). tained, not biased due to dust obscuration or short bursts of massive star formation. To estimate the observed 10 micron flux densities, we used the fits to the HST and FIRES photometry. In Fig. 7 we show the results from a sample with KAB > 25 and 1.5 < zphot < 5. The K-band selection was chosen such the the photometric redshifts are fairly reliable (see Labb´e et al. 2003 for details). The lower limit to the photometric redshift was chosen so that, at an observed wavelength of 10 micron, the SED is still dominated by stellar light and not by a hot dust component. From Fig 7, we conclude that, for the brighter objects, it is possible to obtain good images with signal to noise ratios of 50 within integration times of 25 - 50 hours (see Section 2.). The second important issue is related to the angu- lar size scales of these FIRES objects. In Fig. 8, we plot the effective radius of the objects in FIRES sample versus redshift. The overall trend is that the angular sizes tend to decrease with redshift. Care- ful modelling of the various selection effects involved shows that the physical sizes of luminous galaxies (LV > 2 × 1010Lsun) at 2 < z < 3 are 3 times smaller than that of equally luminous galaxies today (Tru- jillo et al. 2003). The size distribution shows that twice the median effective radius is similar to the resolution (FWHM) of the JWST or in other words, JWST will hardly resolve these distant galaxies. It is clear that for a good study of the morphology of especially the older stellar population, Darwin will be very important. In Rottgering et al. 2000 we have used the models of Devriendt and Guiderdoni (2000) to discuss Darwin's ability to detect and study distant galaxies. It was appropriate to use these models since they include the extinction and emission of dust and as such the models correctly fit the ISO 15 µm source counts. Although a detailed comparison between these mod- els and the FIRES data just presented is in progress, the predicted redshift and angular size distributions also lead to the conclusion that the flux densities and angular sizes are well in the range of Darwin's capa- bilities. 7. CONCLUSION The scientific potential of the Darwin mission for general astrophysical imaging is tremendous. With the present performance of µJy sensitivity and tens of mas angular resolution, it will allow for detailed studies addressing a number of crucial questions re- lated to the formation and evolution of planets, stars, AGN and galaxies. 9 10 micron flux densities estimated from Figure 7. the template fits for a subsample of the FIRES survey with KAB > 25 and 1.5 < zphot < 5. Figure 8. The effective radii as measured in the K- band as a function of redshift for a subsample of the FIRES survey with KAB > 25 ACKNOWLEDGMENTS I would like to acknowledge the pleasant discus- sions and important input from a number of col- leagues including Bruno Guiderdoni, Ewine van Dishoeck, Malcolm Fridlund, Emanuele Daddi, Gi- anLuigi Granato, Marijn Franx, Bjorn Heijligers, Klaus Meisenheimer, Rudolf Le Poole and Jan Willem den Herder. REFERENCES Antonucci R., 1993, ARA&A, 31, 473 Beichman C.A., Woolf N.J., Lindensmith C.A. (eds.), 1999, Terrestrial Planet Finder, Origin of Stars, Planets and Life, Jet Propulsion Laboratory Bell E.F., Baugh C.M., Cole S., Frenk C.S., Lacey C.G., Mar. 2003, ArXiv Astrophysics e-prints Bontemps S., Andr´e P., Kaas A.A., et al., Jun. 2001, A&A, 372, 173 Chapman S.C., Blain A.W., Ivison R.J., Smail I.R., Apr. 2003, Nature, 422, 695 Cole S., Jan. 1991, ApJ, 367, 45 Cole S., Aragon-Salamanca A., Frenk C.S., Navarro J.F., Zepf S.E., Dec. 1994, MNRAS, 271, 781 D'Alessio P., Calvet N., Hartmann L., May 2001, ApJ, 553, 321 D'Arcio L., den Herder J., Le Poole R.S., Rottgering H.J.A., Feb. 2003, In: Interferometry in Space. Edited by Shao, Michael. Proceedings of the SPIE, Volume 4852, pp. 184-195 (2003)., 184 -- 195 Devriendt J.E.G., Guiderdoni B., Nov. 2000, A&A, 363, 851 Efstathiou A., Rowan-Robinson M., Apr. 1995, MN- RAS, 273, 649 Franx M., Moorwood A., Rix H., et al., 2000, The Messenger, 99, 20 Franx M., Labb´e I., Rudnick G., et al., Apr. 2003, ApJ, 587, L79 Fridlund M., et al. M., 2000, Darwin, the Infrared Space Interferometer, Concepts and Feasibility Study Report, ESA report ESA-SCI, 2000, 12 Granato G., Danese L., Franceschini A., Sep. 1997, ApJ, 486, 147 Kauffmann G., White S.D.M., Guiderdoni B., Sep. 1993, MNRAS, 264, 201 Kauffmann G., Guiderdoni B., White S.D.M., Apr. 1994, MNRAS, 267, 981 Krolik J.H., Begelman M.C., Jun. 1988, ApJ, 329, 702 Kurk J., 2003, The Cluster Environment and Gaseous Halos of Distant Radio Galaxies, Ph.D. thesis, , Univ. Leiden, (2003) Labb´e I., Franx M., Rudnick G., et al., Mar. 2003, AJ, 125, 1107 Lacey C., Silk J., Nov. 1991, ApJ, 381, 14 Lacey C., Guiderdoni B., Rocca-Volmerange B., Silk J., Jan. 1993, ApJ, 402, 15 Leger A., Mariotti J.M., Mennesson B., et al., Oct. 1996, Icarus, 123, 249 Manske V., Henning T., Men'shchikov A.B., Mar. 1998, A&A, 331, 52 De Breuck C., Rottgering H., Miley G., van Breugel Monnier J.D., Millan-Gabet R., Nov. 2002, ApJ, 579, W., Best P., Oct. 2000, A&A, 362, 519 694 10 Moorwood A.F., Mar. 1997, In: Proc. SPIE Vol. 2871, p. 1146-1151, Optical Telescopes of Today and Tomorrow, Arne L. Ardeberg; Ed., 1146 -- 1151 Moustakas L.A., Somerville R.S., Sep. 2002, ApJ, 577, 1 Nakajima T., Matsuhara H., Feb. 2001, Appl. Opt., 40, 514 Persi P., Marenzi A.R., Olofsson G., et al., May 2000, A&A, 357, 219 Pier E.A., Krolik J.H., Dec. 1992, ApJ, 401, 99 Rees M.J., 1998, Introductory talk at ngst confer- ence, liege, june 1998 Rees M.J., Ostriker J.P., Jun. 1977, MNRAS, 179, 541 Rosati P., Tozzi P., Giacconi R., et al., Feb. 2002, ApJ, 566, 667 Rottgering H., Granato G., Guiderdoni B., Rudnick G., 2000, In: Lena P.J., Quirrenbach A. (eds.) Pro- ceedings of the SPIE conference Interferometry in Optical Astronomy, Vol. 4006, 742 Rudnick G., Franx M., Rix H., et al., Nov. 2001, AJ, 122, 2205 Rudnick G., Rix H., Franx M., et al., Jul. 2003, ArXiv Astrophysics e-prints Shu F.H., Adams F.C., Lizano S., 1987, ARA&A, 25, 23 Somerville R.S., Primack J.R., Faber S.M., Jan. 2001, MNRAS, 320, 504 Steidel C.C., Adelberger K.L., Giavalisco M., Dick- inson M., Pettini M., Jul. 1999, ApJ, 519, 1 Trujillo I., Rudnick G., Rix H., et al., Jul. 2003, ArXiv Astrophysics e-prints Urry C.M., Padovani P., Sep. 1995, PASP, 107, 803 van Dishoeck E.F., 2000, In: ASP Conf. Ser. 207: Next Generation Space Telescope Science and Technology, 85 Venemans B.P., Kurk J.D., Miley G.K., et al., Apr. 2002, ApJ, 569, L11 White S.D.M., Frenk C.S., Sep. 1991, ApJ, 379, 52 White S.D.M., Rees M.J., May 1978, MNRAS, 183, 341 Wolf S., Henning T., Jan. 1999, A&A, 341, 675
0709.1993
1
0709
2007-09-13T05:40:41
The formation of young B/PS bulges in edge-on barred galaxies
[ "astro-ph" ]
We report about the fact that the stellar population that is born in the gas inflowing towards the central regions can be vertically unstable leading to a B/PS feature remarkably bluer that the surrounding bulge. Using new chemodynamical simulations we show that this young population does not remain as flat as the gaseous nuclear disc and buckles out of the plane to form a new boxy bulge. We show that such a young B/PS bulge can be detected in colour maps.
astro-ph
astro-ph
Formation and Evolution of Galaxy Bulges Proceedings IAU Symposium No. 245, 2008 M. Bureau, E. Athanassoula, and B. Barbuy, eds. c(cid:13) 2008 International Astronomical Union DOI: 00.0000/X000000000000000X The formation of young B/PS bulges in edge-on barred galaxies H. Wozniak1 and L. Michel-Dansac2 1Observatory of Lyon, F-69561 Saint-Genis-Laval cedex, France email: [email protected] 2IATE, Observatorio astronomico, X5000BGR Cordoba, Argentina, email: [email protected] Abstract. We report about the fact that the stellar population that is born in the gas inflowing towards the central regions can be vertically unstable leading to a B/PS feature remarkably bluer that the surrounding bulge . Using new chemodynamical simulations we show that this young population does not remain as flat as the gaseous nuclear disc and buckles out of the plane to form a new boxy bulge. We show that such a young B/PS bulge can be detected in colour maps. Keywords. galaxies: bulges galaxies: evolution galaxies: formation galaxies: nuclei galaxies: stellar content galaxies: structure methods: n-body simulations In the generic case of pure N−body simulations, whenever a disc galaxy forms a bar, a B/PS bulge develops in a few dynamical times. In the case of chemodynamical simula- tions, with gas and star formation/feedback recipes (cf Wozniak & Michel-Dansac (2007), Michel-Dansac & Wozniak (2008) for full details), the B/PS growing process is different and more complicated due to the presence of a young stellar population that is born in the disc. Most of this young population lies in a razor-like central disc during the first 450 Myr, obviously because of the small vertical scaleheight of the initial gas distribu- tion. Since a razor-thin disc is highly unstable the most central part of the disc starts to thicken out the equatorial plane. In roughly a bar rotation period, the vertical distri- bution gets symmetrically peanut shaped over the central 2 kpc, while the young bar is approximately 8 kpc long (cf. Fig 1, t = 600 Myr). At this time, the total mass of the central disc being still low, the thickening process has no clear detectable effect on the whole mass distribution. Then, the old population also starts to buckle out leading to a larger B/PS bulge (cf Fig. 1, t = 1500 Myr). Both populations being fully mixed, the two components can only be splitted in numerical simulations. To observationally detect the young stellar component of a B/PS bulge we have to rely on stellar population tracers, e.g. B−V maps. We thus obtained mock B−V maps calibrat- ing our simulations. The full process is described by Michel-Dansac & Wozniak (2004). We assume that the initial population starts with an age of 10.4 Gyr and has a solar metallicity at z = 0. Due to the colour contrast between both stellar populations, the young B/PS structure is clearly visible (Fig. 2). The dust could however hampered such a detection (Fig. 2, right panels). At t = 600 Myr, since the new stellar disc is young and star formation still active, the dust amount should be likely large. Afterward, the peanut-shape widens out as the young disc evolves. The thick part of the disc doubles its radial size in less than 1 Gyr. The vertical scaleheight also increases with time leading to a well-developed peanut-shaped bulge for t > 1500 Myr. The young B/PS structure, being now much more extended both in the radial and 119 7 0 0 2 p e S 3 1 ] h p - o r t s a [ 1 v 3 9 9 1 . 9 0 7 0 : v i X r a 120 H. Wozniak and L. Michel-Dansac Figure 1. Mass distribution (in log units) of the central 10×5 kpc seen edge-on for T=600 (top panel) and 1500 Myr (bottom). Left panel: all stars (old and young populations) are plotted. Right panel: only the young population has been plotted. Figure 2. B−V colour maps without (left) and with (right) a dust component for the same fields and times. Bluest region are coded in black. The two isocontours have been chosen as to enhance the boxy feature and are quoted at top of each frame. vertical direction, is likely detectable even in the presence of dust. Indeed, the height of the B/PS structure is greater that the dust disc scaleheight. Since the early studies of B/PS bulges, there were many evidences that galaxies hosting such a feature are edge-on barred disc galaxies, and that the B/PS bulges themselves represent the thickest parts of the bars. But stellar populations of B/PS bulges (and their colours) have been rarely studied. Looking for young, blue and small scaleheight B/PS bulge, likely inside older, redder and larger bulge, deserves dedicated surveys. References Michel-Dansac L., Wozniak H., 2004, A&A, 421, 863 Michel-Dansac L., Wozniak H. 2008, A&A submitted Wozniak H., Michel-Dansac L. 2007, MNRAS submitted
astro-ph/0301355
1
0301
2003-01-17T13:31:09
When you wish upon a star: Future developments in astronomical VLBI
[ "astro-ph" ]
In this paper, I present the likely technological development of VLBI, and its impact on the astronomical community over the next 1-5 years. VLBI is currently poised to take advantage of the rapid development in commercial off-the-shelf (COTS) PC-based products. The imminent deployment of disk-based recording systems will enable Gbps data rates to be achieved routinely by both cm and mm-VLBI networks. This, together with anticipated improvements in collecting area, receiver systems and coherence time is set to transform the performance of VLBI in terms of both baseline and image noise sensitivity. At the same time the feasibility of using fibre based communication networks as the basis for production, real-time VLBI networks will begin. Fantastic new correlator output data rates, and the ability to deal with these via powerful PC clusters promises to expand the typical VLBI field-of-view to scales previously reserved for connected, short baseline interferometers. By simultaneously sampling the summed response of all compact radio sources within (and indeed beyond) the half-power point of the VLBI telescope primary beam, simple self-calibration of the target field will ALWAYS be possible at frequencies below a few GHz. Unbiased, broad-band continuum surveys will be conducted over huge areas of sky, and (redshifted) spectral-features will be detected too. By the end of the decade the microJy radio sky will be accessible to VLBI: dozens of sources will be simultaneosuly observed, correlated, detected and fully analysed all within the same day.
astro-ph
astro-ph
New Technologies in VLBI ASP Conference Series, Vol. **VOLUME**, 2003 Y. C. Minh When you wish upon a star: Future developments in astronomical VLBI. M.A. Garrett Joint Institute for VLBI in Europe, Postbus 2, 7990 AA Dwingeloo, The Netherlands Abstract. In this paper, I present the likely technological development of VLBI, and its impact on the astronomical community over the next 1-5 years. VLBI is currently poised to take advantage of the rapid de- velopment in commercial off-the-shelf (COTS) PC-based products. The imminent deployment of disk-based recording systems will enable Gbps data rates to be achieved routinely by both cm and mm-VLBI networks. This, together with anticipated improvements in collecting area, receiver systems and coherence time is set to transform the performance of VLBI in terms of both baseline and image noise sensitivity. At the same time the feasibility of using fibre based communication networks as the ba- sis for production, real-time VLBI networks will begin. Fantastic new correlator output data rates, and the ability to deal with these via pow- erful PC clusters promises to expand the typical VLBI field-of-view to scales previously reserved for connected, short baseline interferometers. By simultaneously sampling the summed response of all compact radio sources within (and indeed beyond) the half-power point of the VLBI telescope primary beam, simple self-calibration of the target field will al- ways be possible at frequencies below a few GHz. Unbiased, broad-band continuum surveys will be conducted over huge areas of sky, and (red- shifted) spectral-features will be detected too. By the end of the decade the microJy radio sky will be accessible to VLBI: dozens of sources will be simultaneosuly observed, correlated, detected and fully analysed all within the same day. 1. Introduction In this paper I review the possible future development of astronomical Very Long Baseline Interferometry (VLBI). My analysis is very much user-driven -- the terms of the review being made via a user-based "wish list". The focus falls on several areas that are important to the astronomical community, and which are likely to be addressed by the new technologies presented throughout this volume. A similar paper is presented by Schluter (this volume) regarding likely advances in Geodetic VLBI activities. Naturally, some of the points addressed here represent areas in which I am personally interested or involved. While these may receive extra attention, I have tried to maintain a reasonable balance throughout the text. The astronomical wish-list thus includes: 1 2 Author & Co-author • vastly improved raw baseline and image sensitivity, • extended frequency coverage & agility, • enhanced image fidelity, • more flexible, robust and reliable VLBI data, including auxiliary (calibra- tion) data, • sharper spatial and spectral resolution, • fantastic new correlator capacities & sufficiently capable offline computing resources (including GRIDs) to analyse the data, • access to a significantly deeper & wider field-of-view, • the introduction of automatic data calibration and new analysis tech- niques. I will review each of these areas in turn, and consider the prospects for possible improvements. It should be noted that this review is made at a cru- cial juncture in the development of VLBI -- in particular the rapid develop- ment of commercial off-the-shelf (COTS) PC-based products are expected to make a substantial impact in many areas of VLBI that have previously enjoyed only incremental advances. These new developments include: flexible digital signal processing, disk-based data recording, data transfer via transnational, broad-band, internet communication networks and capable (offline) computing resources via (Linux) PC clusters. Over the next decade, the opportunities for making substantial progress are excellent. 2. VLBI Sensitivity In VLBI (and radio interferometry in general) the image and baseline sensitivity (σI and σB respectively) are dependent on several different parameters. Two of these are specific to the antenna - the effective collecting area (Ae, m2) and the antenna system noise (Tsys, Kelvin). In addition, sensitivity to broadband con- tinuum radio emission is determined by the output fluctuations of the receiver, and so the signal-to-noise ratio is proportional to the square root of both the spanned observing bandwidth (∆ν Hz) and the total (coherent) integration time (τ , seconds). In SI units the 1σ baseline sensitivity is given by: σB = 2kηb √2∆ντs Tsys1Tsys2 Ae1Ae2 (1) where k is Boltzmanns's constant & ηb accounts for various losses (see Walker et al. 1988 for more details). Note that the baseline sensitivity improves (i.e. σB decreases) proportionally with the geometric average of the effective area of the two telescopes, and inversely proportional to the geometrical average of their system temperatures. APS Conf. Ser. Style 3 Figure 1. Large gains in collecting area can be expected in mm-VLBI with the incorporation of ALMA (top left) and other new mm-capable telescopes (e.g. the KVN top right). The construction of LOFAR (bottom left) is expected to lead to a resurgence in interest in meter- VLBI. On longer times-scales, the SKA (bottom right) will lead (either directly or indirectly) to a substantial increase in the collecting area available to cm-VLBI. 4 Author & Co-author Similarly, the 1σ r.m.s. image noise (assuming optimal data weighting -- natural weighting) is given by: σI = 2kηb √∆ντvuut (Pi 1 Ai Tsysi )2 )2 −Pi( Ai Tsysi (2) It is clear from these expressions that improvements in sensitivity can be obtained via four (largely) independent parameters: (i) increased collecting area, (ii) lower noise receiver systems, (iii) longer coherence times & (iv) larger total observing bandwidths. 2.1. Collecting Area Any increase in collecting area is expected to be incremental at cm wavelengths but over the next decade, significant gains could be made at both higher and lower frequencies (see Figure 1). In particular, mm-VLBI has much to gain by incorporating ALMA within existing networks (see Alef et al. this volume) and similarly low-frequency (meter wavelength) VLBI is bound to be stimulated by the construction of the Low Frequency Array (LOFAR). mm-VLBI (and to a lesser extent cm-VLBI) will also benefit from the addition of new telescopes in Asia (VERA and the KVN -- see Minh et al., Kobayashi et al. and Sassao et al., this volume) and in Europe (the Yebes-OAN 40-m telescope and the IRA 64-m Sardinian Radio Telescope). A significant increase in collecting area for cm-VLBI will have to await the construction of the next generation cm-wave radio telescope -- SKA (see Gurvits this volume). 2.2. Receiver Noise Temperature Both cm and mm-VLBI are likely to see significant progress in the area of improved receiver technology. For example, the e-MERLIN array (see Muxlow et al. this volume) will employ 4-8 GHz receivers that are expected to be a factor of two better than the current (33 K) systems. Even larger factors of improvement can be expected at mm-wavelengths (see Figure 2). 2.3. Coherence Time cm-VLBI Through the application of phase-referencing techniques (Beasley & Conway 1995), it is now possible to extend the coherence time of a cm- VLBI array across the entire duration of the observations. This permits the routine detection of relatively faint (sub-mJy) radio sources (e.g. Garrett et al. 2001) at wavelengths up to λ1 cm. These phase-reference images are often dynamic range limited at the level of 20:1, in-beam phase-referencing removes such limitations and its use (at cm wavelengths) is generally more applicable than usually understood. At λ18 cm for example, in-beam phase-referencing is even possible for very large telescopes, provided wide-field techniques are used to combine the total signal from all compact sources in the primary-beam (see section 7 for more details). This permits in-beam self-calibration to be used in all circumstances (i.e. any target field) -- an important new capability, and one that should be familiar to users of connected arrays. APS Conf. Ser. Style 5 Figure 2. Low-noise amplifiers (such as the broad-band 4-8 GHz OSO/Chalmers system shown left) are set to make significant improve- ments in receivers to be developed for cm but in particular mm and sub- mm wavelengths. Low-noise frequency flexible systems (e.g. the WSRT MFFE system shown right) are now common at cm-wavelengths. mm-VLBI In recent years significant advances have been made in improving the coherence times at mm-wavelengths. VERA for example will use a multi- beam system that simultaneously permits one beam to be directed at the target, and another to a nearby calibrator. Phase corrections from the calibrator are thus continuously applied to the target without any source switching. However, one limitation of dual-beam phase correction at mm wavelengths is that the number of bright mm-calibrators is extremely limited, and thus often involves target-calibrator separations on the scale of several degrees. In these cases, the effects of spatial interpolation errors (across the sky) cannot be ig- nored. Another method which addresses this problem (and one which is showing enormous promise), is to derive mm-wave phase corrections via a cm-wave ref- erence source (see Asaki et al. 1998; also Sasao et al., and Alef et al. this volume). This method relies on the fact that the tropospheric delay is indepen- dent of frequency. In its simplest form, multi-frequency co-axial feed systems can be used to simultaneously observe the target, and phase corrections derived at say 22 GHz are applied (after appropriate scaling with frequency and correct- ing for any antenna based phase-offsets) to much higher frequencies (e.g. 100 GHz). Note that some relative astrometry is also preserved - the position of the cm-wavelength "core" being the reference point. It is clear that this technique might also be usefully employed for high-frequency Space VLBI missions, such as VSOP2 (see Haribyashi et al. this volume). 6 Author & Co-author 2.4. Observing Bandwidth The promise of access to much broader bandwidths is expected to explode in the next decade. The consequences for both mm and cm-VLBI are significant, especially for cm-VLBI where gains in other areas related to sensitivity will be modest. Currently 256 Mbps is just about the maximum data rate that can be currently sustained in most VLBI networks. For the European VLBI Network (EVN, see www.evlbi.org) the limitation is not a technological one (512 Mbps recording is now routinely performed) but the availability of thin-tapes. Within the next 2 years sustained data rates of 1 Gbps will certainly be available to VLBI users. This will lead to a factor of two better sensitivity for both mm amd cm-VLBI networks. The longer term aim must be to attain data rates of several Gbps, reaching tens of Gbps by the end of the decade. This can be utilised by both mm and cm-VLBI networks, in the latter case not just for bandwidth but in order to employ multi-bit signal representation required by RFI mitigation algorithms. PC disk-based recorders The promise of access to much broader bandwidths is expected to explode in the next decade (see Figure 3). The consequences for both mm and cm-VLBI are significant, especially for cm-VLBI where gains in other areas will be modest. The maximum total bandwidth currently used in VLBI is ∼ 64 MHz (in each of 2 polarisations), corresponding to a total data rate of 512 Mbps (2-bit signal representation and Nyquist Sampling). The replacement of the current generation of magnetic tape recorders, with PC-based disk recorder systems (see Whitney et al., Parsley et al., Romney et al., Kondo et al. this volume) is expected to take place over the next few years. By employing commercial PC hardware, the VLBI community will be able to take advantage of the rapid technological development in this area. In principle, a doubling of the data capacity of PC-based recorders might be expected every few years. Since the Mk5A system can already record at 1 Gbps (see Figure 3), data rates in excess of this are likely to be possible on relatively short time-scales. In addition, the cost of disks will continue to shrink, at least in real terms (i.e. for a given storage capacity). Real Time, Optical Fibre-based VLBI networks An alternative (or perhaps suc- cessor) to disk-based recorder systems is the connection of VLBI networks via optical fibres (see Parsley & Whitney this volume & Figure 3). Fibre communi- cation networks are ideally suited to the real-time transfer of huge amounts of data over long distances. The adoption of direct fibre connections by e-MERLIN (see Muxlow this volume) signals the progress that is being made in this area. As the costs of these networks continue to fall, and as commercial networks be- come more flexible, the introduction of a real-time VLBI system is a reasonable goal to pursue. In Europe there are plans to demonstrate the feasibility of real-time VLBI using shared IP routed networks (e.g. G´EANT). A proof-of-concept test pro- gramme (to be conducted over the next 1-2 years), aims to connect together directly, at least 4 European telescopes to the EVN correlator at JIVE (see Parsley this volume). Each telescope will generate up to 1 Gbps data streams, APS Conf. Ser. Style 7 and these will feed into the EVN correlator at JIVE (see Parsley et al. this vol- ume). The idea is to correlate the data with minimal buffering at either end of the fibres. The provision of local loops (last mile connections) to the telescopes and correlator is the critical item. Local loops are in place at Dwingeloo (JIVE) and Torun. Westerbork is to be connected in mid-2003, and negotiations are on-going at other EVN sites, in particular Jodrell Bank, Effelsberg, Medicina, Onsala and Metsahovi. Technical issues that need to be thrashed out include the quality of service required by VLBI, and the amount of buffering required at the telescopes and correlator. Assuming the first tests are successful, the ambition in Europe is to investigate "production" real-time VLBI networks, and to broaden participation to include Asia, North-America and Africa. Implications for VLBI Back/Front-end systems & Correlators Both fibre and PC-based VLBI networks will permit routine Gbps VLBI observations to be made in the course of the next few years. Developments beyond this requires (in many cases) a replacement to the current VLBI data acquisition system (in particular the expensive, and now obsolete, analogue Base Band Converters - BBCs) with cheaper and more flexible digital replacement systems. The interest in the latter topic is witnessed by the activity reported in this volume (see Ferris et al., Tuccari et al., Ying et al., Roh et al., Kondo et al., Koyama et al., Iguchi et al). However, discussion about the necessity to broad-band front-end telescope receiver systems was limited. Correlation of Gbps data stream is also a problem. For example, the EVN MkIV correlator at JIVE can currently handle 16 telescopes at 1 Gbps or (po- tentially) 8 telescopes at 2 Gbps. Data rates in excess of this would require a new, more capable correlator, similar to that being developed for the EVLA and e-MERLIN (see Carlson et al. this volume). It is clear that to take full advantage of the increasing capacity of both disk and fibre-based systems, considerable efforts must begin now, in terms of new receiver systems, replacement back-end data acquisition racks and future VLBI correlator developments. 2.5. Overall Sensitivity Gains & Image Fidelity It is clear that both cm-VLBI and in particular mm-VLBI can expect to make considerable gains in terms of collecting area, bandwidth (data rates), receiver noise temperature and techniques to extend the coherence time of the data. For continuum cm-VLBI a total gain in sensitivity of at least 5 seems plau- sible over the next few years. In principle, noise levels at microJy and even sub-microJy levels should be attainable (see Figure 4) by Global VLBI arrays. An important provision is that both the EVN and Very Long Baseline Array (VLBA, see www.nrao.edu) adopt the same fully compatible, next generation (disk-based) data acquisition systems. It is good to see that the Global VLBI Working Group (that also met here in Korea) have already started to worry about these very issues. In the case of mm-VLBI, at least an order of magnitude improvement would not be surprising. In addition, since mm-VLBI's gains will also include addi- 8 Author & Co-author Figure 3. Magnetic tape technology (top left) is now being replaced by PC disk-based (Mk5) recording systems (top right). Real-time VLBI using optical fibre networks must be the long-term goal (bottom right). The EVN correlator at JIVE is already connected by a fibre network that currently provide Gbps data rates but will be easily upgraded to permit even larger data rates to be employed. APS Conf. Ser. Style 9 Figure 4. bandwidth and integration time. The sub-microJy sensitivity of the e-EVN as a function of tional collecting area and improved receiver systems, spectral-line studies also stand to benefit, not just standard continuum observations. So far we have neglected to mention that any increase in observed band- width will also (assuming the data remains unaveraged in frequency) result in an improvement in uv-coverage and thus image fidelity. This is particularly the case for cm-VLBI where the fractional bandwidth should soon approach unity and Multi Frequency Synthesis (MFS) techniques can be employed to take full advantage of this. Figure 5 presents the uv-coverage of the e-EVN assuming a total bandwidth of 2 GHz per polarisation. 3. Frequency Coverage/Flexibility, Receiver Design & RFI An important development in recent times, has been the construction of front- end receivers that can be instantaneously tuned over a wide-range of sky fre- quency (e.g. from 1-10 GHz in the case of the Allen Telescope Array). In these systems, several bands (each ∼ 1 GHz in extent) can then be selected individu- ally and digitised. Systems like these need to be in place at VLBI telescopes if we are to fully capitalise on the expected increase in capacity of second genera- tion disk and (first generation) fibre-based VLBI networks. However, in addition to sensitivity and image fidelity issues (see previous section), there is another astronomical motivation for instantaneous access to large swathes of bandwidth: Serendipitous VLBI spectral line surveys (e.g. HI in absorption). Already such surveys are being conducted by connected element arrays (e.g. Morganti & Gar- rett 2002) and as the field-of-view of VLBI observations increases (see section 6), VLBI can easily follow suit. 10 Author & Co-author Figure 5. Left: The current uv-coverage of the EVN at λ6 cm for a source located at δ = 30◦. Right: the extended (almost full) uv- coverage of the eEVN for the same source, assuming a total bandwidth of 2 GHz per polarisation. The effects of Radio Frequency Interference (RFI) will also become increas- ingly important as VLBI systems become more sensitive and observe larger bandwidths. Although RFI does not usually correlate on baselines > 10 km, lo- cal interfering signals are often so strong that they can easily saturate receivers, and dominate the antenna system noise. In addition, as a noise source, RFI is often extremely variable on time scales of a few seconds or less -- tracking the telescopes calibration under such conditions is usually impossible. Real VLBI users are often sceptical of many RFI suppression techniques but at this meeting there were several good presentations that suggest there are more effective ways to counter RFI (e.g. Kesteven and Roshi this volume) than simply deleting the data. We had better start using these sooner, rather than later. Rapid frequency switching is a routine observing mode for the VLBA. This capability is important for spectral index mapping studies but also increases the robustness and reliability of VLBI operations. Many of the telescopes in the EVN are now frequency flexible (see Figure 2) but only a few experiments have taken advantage of this facility so far. Rapid progress is expected to take place in this area over the next year. Finally, polarisation purity is another topic that often gets ignored in VLBI receiver design. This is another area in which a homogeneous array such as the VLBA has a significant advantage. The current aim of the EVN is to produce < 2% cross-talk between left and right hand circular polarisation channels. Al- though many of the EVN telescopes are actually much better than this, some telescopes show cross-talk at the level of ∼ 15%! These include special cases such as the WSRT phased-array. These figures can limit the dynamic range of total intensity images of even moderately bright radio sources, and for polarisa- tion studies, the level of impurity is large enough that second order calibration corrections (usually assumed to be negligible) must be accounted for. APS Conf. Ser. Style 11 4. Sharper Resolution, Space VLBI and SKA Configurations When it comes to resolution VLBI astronomers are a difficult lot to satisfy. Their desire for increasingly better resolution forces them to move towards higher frequencies and/or longer baselines. This is an effect that was clearly demonstrated at this meeting. In particular, next generation orbiting VLBI telescopes (e.g. VSOP-2) combine both these elements together, employing only high-frequency receivers (8, 22 & 43 GHz), and baselines between 3 − 5 Earth radii (see Hirabayashi et al. and Mochizuki et al. Gurvits et al. this volume). RadioAstron goes a step further with even longer baselines being proposed. I feel it necessary to introduce a note of caution at this point. As we move towards higher frequencies and longer baselines, AGN science clearly stands to benefit. However, other areas of growing importance are being neglected e.g. the study of SNe, SNR, micro-quasars, active stars, HI absorption, OH emission, starburst galaxies, high-z star-forming/AGN systems etc. These require high resolution too - perhaps third generation Space VLBI missions will be able to address these requirements too. Angular resolution is also a hot topic in the discussion of array configura- tions for the SKA (see Gurvits this volume). Plans for the next generation of ground and space based astronomical observatories, will provide much higher resolution (approaching traditional milliarcsecond scales) for sub-mm, IR, op- tical/UV and x-ray astronomers. This, together with source confusion is likely to see the SKA deployed with complimentary baselines in excess of 1000 km (Garrett et al. 2002). 5. VLBI Calibration, auxiliary data products and automatic data (pipeline) analysis Over the last 10 years the VLBA has the set the standard in terms of generating accurate and homogeneous astronomical VLBI calibration data. Meanwhile the rest of us have been playing "catch-up"! The generation of such data is vital in order to make VLBI transparent to all astronomers (not just a few "black-belt" practitioners). In addition, it is necessary in order to obtain high-dynamic range images via both manual and automatic analysis paths (but especially the latter). The EVN is beginning to get there - continuous system temperature data is now available, both as a function of time and frequency (see Figure 6). The calibration of the telescopes (via the NASA/GSFC Field System) is considerably improved, with lots of essential new features (see Himwich this volume). In addition, the EVN is now able to compare the pointing position of the telescope and the direction of the target source i.e. it is now possible to generate "flag files" that can be used to identify non-valid telescope data (see also Figure 6). Progress in this area was also reported by other arrays (e.g. the LBA, Tingay this volume). As data sets become larger (see section 6 & 7) it will be necessary to au- tomatically perform on-line calibration and data analysis. The EVN pipeline (Reynolds, Paragi & Garrett 2002) is the first step along this road. All EVN projects are now "pipelined" by default. The products include a set of AIPS calibration tables (a-priori calibration, fringe-fitting and self-calibration) and 12 Author & Co-author Figure 6. In the 1990's the VLBA set the standard for auxiliary data products. Other VLBI networks are only now catching up. Continuous TSYS (left) and the application of Telescope Flag files (right) are now routinely generated by the EVN. various standard plots. As well as reducing the effort required on behalf of the astronomer, we also gain a much better understanding of the performance of the network. The default mode is only to pipeline data associated with calibrators but on request the target source can also be analysed. All astronomers can take advantage of the EVN pipeline - irrespecitve of their experience, affiliation or geographical location. 6. Field-of-View, Spectral-line & Fantastic Correlator output data rates For a connected element array, the field-of-view is often set by the primary beam size of the individual telescope elements. For VLBI this is hardly ever the case. In VLBI, a more demanding limitation is set by the fine spectral resolution and short integration times that must be employed in order to circumvent both band- width smearing and time averaging effects. Since preserving the field-of-view scales (computationally) with baseline length squared, wide-field VLBI analysis places enormous pressure on offline computer resources (processing speed and disk space). These are many orders of magnitude greater than for short-baseline, connected arrays. In the same vein, a VLBI field-of-view that is comparable with the primary beam of the individual telescopes, places demands on the correlator output data rate that are nothing short of "fantastic". Pushed to their limits, current VLBI correlators are just about capable of providing sufficient resolution in both time and frequency (e.g. 0.5 sec integration time and 1024 × 62.5 kHz channels) to permit the inner 3 arcmin the telescope primary beam to be imaged out with full sensitivity at 1.4 GHz. This corresponds to a correlator output data rate (∼ 1 Mbyte/sec) - well short of what is required to map-out the full (half- power point) primary beam of a 100-m, never-mind the much larger field-of-view associated with 25 or 32-m class telescopes. The PCInt project currently being developed at JIVE (see Figure 7), will enable the EVN correlator to generate and handle output data rates as high APS Conf. Ser. Style 13 Figure 7. The EVN correlator at JIVE has been operational since the end of 1999. The PCInt project will enable the EVN correlator to generate and handle output data rates as high as 160 MBytes/sec or 13 TBytes/day. This will permit the full capacity of the correlator to be harnessed. as 160 MBytes/sec or 13 TBytes/day. This will permit the full capacity of the correlator to be harnessed, permitting a spectral resolution of 8092 channels per baseline or integration times as short as 15 milli-seconds. PCInt will hugely expand the field-of-view of VLBI quite generally (not just the EVN), and will provide the capability to simultaneously map-out large swathes of the radio sky with milliarcsecond resolution. Much finer spectral resolution is not only required to expand the field-of- view but it is also important for spectral-line studies. Often spectral-line projects have to trade spectral resolution for the number of telescopes, polarisation prod- ucts etc. Many projects also require multiple-pass correlation because more than one spectral feature is present (and the correlator provides just enough resolu- tion for one line in any given pass). In addition, broad-banding of connected interferometers (in particular the upgraded WSRT) has recently revealed some very broad HI absorption systems (e.g. Morganti et al. 2002), broad enough, that current VLBI correlators are inadequate to appropriately sample the full width of the line. 7. Deep, Wide-Field cm-VLBI Studies The application of wide-field techniques to VLBI data analysis is fundamental to high resolution, deep field studies of the faint sub-mJy radio source population. That VLBI can make a contribution in this area was first demonstrated with the EVN 1.6 GHz observations of the Hubble Deep Field (HDF). 7.1. EVN Observations of the HDF EVN 1.6 GHz observations of the HDF (Garrett et al. 2001) were the first VLBI observations of what is essentially a "blank field", i.e. a region of sky, devoid of bright radio sources. The brightest source in the HDF-N (as measured by the WSRT and VLA) is a 1.6 mJy FR1 radio galaxy at z = 1. In addition, the 14 Author & Co-author observations (correlated at the VLBA correlator in Socorro) employed wide-field techniques (1 sec integrations, 64 × 125 kHz channels) and it was thus possible to simultaneously image out the full field encompassed by the HDF-N (∼ 6 sq. arcmin). Three sources were detected within this field (see Figure 8) - including the faintest source yet detected by VLBI - a 180 microJy AGN associated with a spiral galaxy (with a bulge) at z = 0.96. EVN detections in the HDF: the distant z=1.01 FRI (top), Figure 8. the z=4.4 dusty obscured starburst hosting a hidden AGN (middle) and the faint 180 microJy, z=0.96 AGN (bottom). Crosses represent the MERLIN-VLA positions for these sources. 7.2. Recent VLBA+GBT deep field results The rms noise levels achieved by the EVN HDF-N observations were limited by phase errors introduced via conventional, external phase-referencing (switching) techniques. The field-of-view was limited by the frequency and time resolution that could then be achieved by the VLBA correlator. APS Conf. Ser. Style 15 Some recent VLBA+GBT deep field observations illustrate the gains to be made in employing "in-beam" phase referencing. Figure 9 shows the deepest VLBI images made to date (Garrett, Wrobel & Morganti in prep). The images (with an rms noise of 9 microJy/beam in the centre of the field) were made from a 1.4 GHz VLBA+GBT observing run (3 × 8 hours @256 Mbps). In-beam phase-referencing was used to provide essentially perfect phase corrections for this data set, and eight sources are simultaneously detected (> 7σ) within and outside the half-power point of the GBT primary beam. Of these eight sources, two sub-mJy sources are detected within the primary beam of the GBT, in addition to the in-beam phase reference calibrator (a compact 20 mJy source, first detected by Wrobel et al.). The images of sources far from the field centred are tapered, since the time/frequency sampling is only adequate for sources that lie within the primary beam of the GBT (the latter being centred on the VLBI phase centre). The total (target) data set size is 60 Gbytes (0.5 secs integration, 1024 × 62.5 kHz channels). Images were made with the AIPS task IMAGR - dirty maps/beams of each sub-band (IF) for each epoch were generated blindly and then simply co-added together. Each postage stamp image took about 8 hours to produce on a dual (2 GHz) processor Linux box. The analysis of these data is on-going. For sources that were bright enough, CLEAN maps were produced by simply subtracting the dirty beam from the dirty image (AIPS task APCLN). More complicated tasks (e.g. IMAGR) involving a visibility based CLEAN are prohibitively expensive in terms of CPU requirements. 7.3. Lessons for the future Phase-referencing no longer required at 1-3 GHz What is clear now, and what I would like to focus on in the final section of this paper, is that if the field of view of VLBI can indeed be expanded (see section 6 and the PCInt development at JIVE) then the whole concept of phase-referencing has to change - at least at frequencies below 3 GHz. The reason for this is that if one can simultaneously sample the summed response of all compact radio sources within (and indeed beyond) the half-power point of the VLBI telescope primary beam, then simple self-calibration of the target field is always possible, essentially trivial, and can provide (essentially) perfect phase corrections to the data. Phase-referencing is still required (to some extent) in order to preserve the astrometry and to improve the coherence time of the data - before self-calibration of the target field is attempted. In simple terms, at frequencies 1-3 GHz (perhaps even higher or lower) a VLBI observation (correlated and analysed using wide-field techniques) is pretty well like any connected element array data set! Phase stability and coherence times are essentially infinite. From a technical perspective the message is clear: (i) With sustained data rates of 1 Gbps, Global VLBI can approach, in some cases surpass, the rms noise levels attained by connected element arrays and (ii) every VLBI target field can be self-calibrated at frequencies ∼ 1 − 3 GHz, provided a wide enough field of view can be maintained, and sufficient computing resources are available to cope with the enormous data sets implied. 16 Author & Co-author Figure 9. Deep VLBA+GBT 1.4 GHz observations of a small portion of the NOAO-N Bootes deep field. The VLBI detections are shown inset. Radio line contours (produced by the WSRT) are superimposed on the NOAO optical field). One non-detection is also shown (bottom left) - a bright, presumably nearby spiral galaxy that is well detected by the WSRT (around the few mJy level). Very likely the radio emission from this system is associated with extended star formation. These are the deepest images made with VLBI to date (Garrett, Wrobel & Morganti in preparation). APS Conf. Ser. Style 17 Once systems like PCInt become available, the bottleneck quickly moves towards the problem of handling the enormous data sets that such systems can generate. Clusters of Linux PCs are the only feasible solution today, together with massive storage devices. Access to GRID like computing resources may be the only feasible short-term solution. 8. A personal view of the future The sensitivity of both mm and cm-VLBI is set to improve dramatically. In the next 5-10 years, Gbps data rates will be routine, coherence times will be virtually unlimited and the first real-time eVLBI production networks will begin to be realised. As network performance is continuously monitored, serious telescope failures will be rare, and feedback immediate. Target of Opportunity observa- tions (e.g. GRBs) will be possible even for part-time arrays such as the EVN. Frequency bands will be configurable and blind searches for spectral features in the data will be possible. Wide-field imaging will be the norm, and several dozens of sources will be detected and imaged simultaneously, in any given ob- serving run. Huge VLBI source surveys will be conducted without relying on the biased selection criteria employed in major surveys today. The resulting wide- field data sets will be usefully mined by Virtual Observatory facilities, providing on-the-fly images and spectra of particular regions of sky. The whole process of doing VLBI will be irevokobaly changed, and the scientific base of our observations expanded immesureably. We can look forward to an era in which, for the first time, VLBI observations will be made, correlated and automatically pipelined all within the same day! I'd like to thank Young Chol Minh and the rest of the LOC for their hospitality and congratuate them on organising and hosting an excellent meeting. 9. References Asaki et al. 1998, Radio Science, 33, 1297. Beasley, A. J., & Conway, J. E. 1995, ''VLBI " Phase-Referencing,'' in ASP Conf. Series 82, Very Long Baseline Interferometry and the VLBA, ed. J. A. Zensus, P. J. Diamond, & P. J. Napier (San Francisco: ASP), 327-343. Garrett, M.A., Muxlow, T.W.B., Garrington, S.T. et al. 2001, A&A Letters, 366, L5 (\protect\vrule width0pt\protect\href{http://arxiv.org/abs/astro-ph/0008509}{astro-ph/0008509}). Garrett, M.A. 2002, in Procs. of the 6th EVN Symposium, ed. Ros, E. et al. (\protect\vrule width0pt\protect\href{http://arxiv.org/abs/astro-ph/0211013}{astro-ph/0211013}). Morganti, R. & Garrett, 2002, M.A. WSRT Newsletter No. 17, p6. Morganti, R. et al. 2002, in Proceedings of the 3rd GPS/CSS Workshop, ed. T. Tzioumis, W.de Vries, I. Snellen, & A. Koekemoer (\protect\vrule width0pt\protect\href{http://arxiv.org/abs/astro-ph/0212321}{astro-ph/0212321}). 18 Author & Co-author Reynolds, R. Paragi, Z. \& Garrett, M.A. 2002, Presented at the URSI General Assembly (\protect\vrule width0pt\protect\href{http://arxiv.org/abs/astro-ph/0205118}{astro-ph/0205118}). Walker, R.C., 1988, in VLBI Techniques & Applications ed. Felli & Spencer, Kluwer Academic Publishers.
astro-ph/0007277
1
0007
2000-07-19T13:17:12
Scattering of Dirac Waves off Kerr Black Holes
[ "astro-ph", "gr-qc" ]
Chandrasekhar separated the Dirac equation for spinning and massive particles in Kerr geometry into radial and angular parts. Here we solve the complete wave equation and find out how the Dirac wave scatters off Kerr black holes. The eigenfunctions, eigenvalues and reflection and transmission co-efficients are computed. We compare the solutions with several parameters to show how a spinning black hole distinguishes mass and energy of incoming waves. Very close to the horizon the solutions become independent of the particle parameters indicating an universality of the behaviour.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 31 October 2018 (MN LATEX style file v1.4) Scattering of Dirac Waves off Kerr Black Holes S. K. Chakrabarti and Banibrata Mukhopadhyay S.N. Bose National Centre for Basic Sciences, JD-Block, Sector III, Salt Lake, Calcutta 700091, India [email protected] and [email protected] Accepted for publication in MNRAS ABSTRACT Chandrasekhar separated the Dirac equation for spinning and massive particles in Kerr geometry into radial and angular parts. Here we solve the complete wave equation and find out how the Dirac wave scatters off Kerr black holes. The eigenfunctions, eigenvalues and reflection and transmission co-efficients are computed. We compare the solutions with several parameters to show how a spinning black hole distinguishes mass and energy of incoming waves. Very close to the horizon the solutions become independent of the particle parameters indicating an universality of the behaviour. Key words: Black holes -- spin-1/2 particles -- Dirac equations -- Waves: scattering 1 INTRODUCTION Chandrasekhar (1976) separated Dirac equation in Kerr black hole geometry into radial 0 0 0 2 l u J 9 1 1 v 7 7 2 7 0 0 0 / h p - o r t s a : v i X r a (r) and angular (θ) parts. The radial equations governing the radial wave-functions R± 1 corresponding to spin ± 1 2 are given by (with -- h = 1 = G = c): 2 ∆ 1 2D0R− 1 2 = ( -- λ + impr)∆ 1 2 R+ 1 2 ; ∆ 1 2D† 0∆ 1 2 R+ 1 2 = ( -- λ − impr)R− 1 2 , where, the operators Dn and D† n are given by, iK ∆ + 2n (r − M) ∆ ; D† n = ∂r − iK ∆ + 2n (r − M) ∆ , Dn = ∂r + c(cid:13) 0000 RAS (1) (2) S. K. Chakrabarti and Banibrata Mukhopadhyay 2 and ∆ = r2 + a2 − 2Mr; K = (r2 + a2)σ + am. (3) Here, a is the Kerr parameter, n is an integer, σ is the frequency of incident wave, M is the mass of the black hole, mp is the rest mass of the Dirac particle, -- λ is the eigenvalue which is the separation constant of complete Dirac equation and m is the azimuthal quantum number. The equations governing the angular wave-functions S± 1 2 corresponding to spin ± 1 2 are given by: L 1 2 S+ 1 2 = −( -- λ − amp cos θ)S− 1 2 ; L† 1 2 S− 1 2 = +( -- λ + amp cos θ)S+ 1 2 where, the operators Ln and L† n are given by, Ln = ∂θ + Q + n cot θ; L† n = ∂θ − Q + n cot θ and Q = aσ sin θ + m cosec θ. (4) (5) (6) Combining eqs. 4, one obtains a second order angular eigenvalue equations which admits exact solutions for spin-half particles when ρ = mp (7) (8) -- λ2 = (l + 1 2 )2 + aσ(p + 2m) + a2σ2"1 − σ = 1 (Chakrabarti, 1984), 2(l + 1) + aσx# , y2 and where, and Slm = 1 2 1 2 Ylm − aσy 2(l + 1) + aσx Yl+1m 1 2 p = F (l, l); x = F (l + 1, l + 1); y = F (l, l + 1) F (l1, l2) = [(2l2+1)(2l1+1)] 1 2 < l21m0l1m > [< l21 [< l21 1 0l1 2 > +(−1)l2−lρ√2 < l21− 1 > +(−1)l2−l < l21m0l1m > (9) 1 2 >]. 2 1l1 1 1 2 1 2 20l1 Here, < ...... > are the usual Clebsh-Gordon coefficients and sYlm are the standard spin- weighted spherical harmonics (Chakrabarti, 1984; see also, Goldberg et al 1967, Breure et al, 1982) of spin s and usual quantum numbers l and m. When ρ 6= mp σ = 1, one obtains the c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Scattering of Dirac Waves off Kerr Black Holes 3 solutions perturbatively with aσ to be the perturbation parameter. The detailed procedure to obtain eigenfunctions and eigenvalues is in (Chakrabarti 1984) and is not described here. The radial equations (1) are in coupled form. One can decouple them and express the equation either in terms of spin up or spin down wave functions R± 1 2 but the expression loses its transparency. It is thus advisable to use the approach of Chandrasekhar (1983) by changing the basis and independent variable r to, r∗ = r + where, 2Mr+ + am/σ r+ − r− log r r+ − 1! − 2Mr− + am/σ r+ − r− log r r− − 1! (r > r+). d dr∗ = ∆ ω2 d dr ; ω2 = r2 + α2; α2 = a2 + am/σ, (10) (11) to transform the set of coupled equations (eq. 1) into two independent one dimensional wave equations given by: = dr∗ − iσ! P+ 1 d ∆ ( d 2 Here, D0 = ω2 as R− 1 = P− 1 dr∗ and ∆ 2 2 1 2 ∆ ω2 ( -- λ − impr)P− 1 ∆ ( d 0 = ω2 2 + iσ) and D† 1 2 R+ 1 2 = P+ 1 2 . ; d dr∗ + iσ! P− 1 2 = 1 2 ∆ ω2 ( -- λ + impr)P+ 1 2 . (12) dr∗ − iσ) were used and wave functions were redefined 2 SOLUTION PROCEDURE We define a new variable, θ = tan−1(mpr/ -- λ), which gives, ( -- λ ± impr) = exp(±iθ)√( -- λ2 + m2 pr2). (13) Also define, P+ 1 2 = ψ+ 1 2 Z± = ψ+ 1 2 ± ψ− 1 2 and further choosing r∗ = r∗ + 1 1 2 exp(cid:20)− i tan−1(cid:18) mpr -- λ (cid:19)(cid:21) ; 2σ tan−1( mpr -- λ the above equations become, P− 1 2 = ψ− 1 2 exp(cid:20)+ 1 2 i tan−1(cid:18)mpr ) so that dr∗ = (1 + ∆ ω2 (14) -- λ (cid:19)(cid:21) , pr2 )dr∗, and +m2 1 -- λmp 2σ -- λ2 dr∗ − W! Z+ = iσZ−; d d dr∗ + W! Z− = iσZ+, where, W = ∆ 1 2 ( -- λ2 + m2 pr2)3/2 ω2( -- λ2 + m2 pr2) + -- λmp∆/2σ . ¿From these equations, we readily obtain a pair of independent one-dimensional wave equa- tions, d2 dr∗ 2 + σ2! Z± = V±Z±; where V± = W 2 ± dW dr∗ . (17) c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 (15) (16) 4 S. K. Chakrabarti and Banibrata Mukhopadhyay By transformation of the variable from r to r∗ (and r∗) the horizon is shifted from r = r+ to r∗ = −∞ unless σ ≤ σs = −am/2Mr+ (eq. 10). In this connection, it is customary to define σc where α2 = 0 (eq. 11). Thus, σc = −m/a. If σ ≤ σs, super-radiation is expected for particles with integral spins but not for those with half-integral spin (Chandrasekhar 1983). Thus, we concentrate on the region where, σ > σs. The choice of parameters is generally made in such a way that there is a significant interaction between the particle and the black hole, i.e., when the Compton wavelength of the incoming wave is of the same order as the outer horizon of the Kerr black hole. Similarly, the frequency of the incoming particle (or wave) should be of the same order as inverse of light crossing time of the radius of the black hole. These yield, mp ∼ σ ∼ [M +q(M 2 − a2)]−1. (18) Thus, we need to deal with quantum black holes to get 'interesting' results. There are two cases of interest: (1) the waves do not 'hit' the potential barrier and (2) the waves do hit the potential barrier. First, we replace the potential barrier by a large number of steps as in the step-barrier problem in quantum mechanics. Fig. 1 shows one such example of the potential barrier V+ (eq. 17) which is drawn for a = 0.5, mp = 0.8 and σ = 0.8. In reality we use tens of thousands of steps with suitable variable widths so that the steps become indistinguishable from the actual function. The solution of (17) at nth step can be written as (Davydov 1976), when energy of the wave is greater than the height of the potential barrier. The standard Z+,n = Anexp[iknr∗,n] + Bnexp[−ikn r∗,n] (19) junction condition is given as (Davydov 1976), Z+,n = Z+,n+1 and dZ+ dr∗ n = dZ+ dr∗ n+1. The reflection and transmission co-efficients at nth junction are given by: (20) (21) Rn = An+1(kn+1 − kn) + Bn+1(kn+1 + kn) An+1(kn+1 + kn) + Bn+1(kn+1 − kn) ; Tn = 1 − Rn At each of the n steps these conditions were used to connect solutions at successive steps. Here, k is the wave number (k = qσ2 − V±) of the wave and kn is its value at nth step. We use the 'no-reflection' inner boundary condition: R → 0 at r∗ → −∞. For the cases where waves hit on the potential barrier, inside the barrier (where σ2 < V+) c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Scattering of Dirac Waves off Kerr Black Holes 5 Fig. 1 : Behaviour of V+ (smooth solid curve) for a = 0.5, mp = 0.8, σ = 0.8. This is approximated as a collection of steps. In realily tens of thousand steps were used with varying step size which mimic the potential with arbitrary accuracy. we use the wave function of the form Z+,n = Anexp[−αn r∗,n] + Bnexp[αnr∗,n] (22) where, αn = qV± − σ2, as in usual quantum mechanics. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 6 S. K. Chakrabarti and Banibrata Mukhopadhyay Fig. 2a: Reflection (R) and transmission (T ) coefficients of waves with varying mass as functions of r∗. mp = 0.78 (solid), mp = 0.79 (dotted) and mp = 0.80 (long-dashed) are used. Other parameters are a = 0.5 and σ = 0.8. Inset shows R in logarithmic scale which falls off exponentially just outside the horizon. 3 EXAMPLES OF SOLUTIONS Fig. 2a shows three solutions [amplitudes of Re(Z+)] for parameters: a = 0.5, σ = 0.8 and mp = 0.78, 0.79, and 0.80 respectively in solid, dotted and long-dashed curves. The energy σ2 is always higher compared to the height of the potential barrier (Fig. 1) and therefore the particles do not 'hit' the barrier. k goes up and therefore the wavelength goes down c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Scattering of Dirac Waves off Kerr Black Holes 7 Fig. 2b: Amplitude of Re(Z+) of waves with varying mass as functions of r∗. mp = 0.78 (solid), mp = 0.79 (dotted) and mp = 0.80 (long-dashed) are used. Other parameters are a = 0.5 and σ = 0.8. monotonically as the wave approaches a black hole. It is to be noted that though ours is apparantly a 'crude' method, it has flexibility and is capable of presenting insight into the problem, surpassing any other method such as ODE solver packages. This is because one can choose (a) variable steps depending on steepness of the potential to ensure uniform accuracy, and at the same time (b) virtually infinite number of steps to follow the potential c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 8 S. K. Chakrabarti and Banibrata Mukhopadhyay as closely as possible. For instance, in the inset, we show R in logarithmic scale very close to the horizon. All the three curves marge, indicating that the solutions are independent of the mass of the particle and a closer inspection shows that here, the slope of the curve depends only on σ. The exponential dependence of Rn close to the horizon becomes obvious. Asymptotically, V± = m2 p (eq. 17), thus, as mp goes down, the wavelength goes down. In Fig. 2b, we present the instantaneous values of the reflection R and transmission T coefficients (i.e., Rn and Tn of eq. 21) for the same three cases. As the particle mass is decreased, k goes up and corresponding R goes down consistent with the limit that as k → ∞, there would be no reflection at all as in a quantum mechanical problem. Figs. 3(a-b) compare a few solutions where the incoming particles 'hit' the potential barrier. We choose, a = 0.95, σ = 0.168 and mass of the particle mp = 0.16, 0.164, 0.168 respectively in solid, dotted and long-dashed curves. Inside the barrier, the wave decays before coming back to a sinusoidal behaviour, before entering into a black hole. In Fig. 3b, we plotted the potential (shifted by 2.05 along vertical axis for clarity). Here too, the reflection coefficient goes down as k goes up consistent with the classical result that as the barrier height goes up more and more, reflection is taking place strongly. Note however, that the reflection is close to a hundred percent. Tunneling causes only a few percent to be lost into the black hole. Figs. 4(a-b) show the nature of the complete wave function when both the radial and the angular solutions (Chakrabarti 1984) are included. Fig. 4a shows contours of constant amplitude of the wave (R−1/2S−1/2) in the meridional plane -- X is along radial direction in the equatorial plane and Y is along the vertical direction. The parameters are a = 0.5, mp = 0.8 and σ = 0.8. Some levels are marked. Two successive contours have amplitude difference of 0.1. In Fig. 4b a three-dimensional nature of the complete solution is given. Both of these figures clearly show how the wavelength varies with distance. Amplitude of the spherical wave coming from a large distance also gets weaker along the vertical axis and the wave is forced to fall generally along the equatorial plane, possibly due to the dragging of the inertial frame. 4 CONCLUSION Scatterings of massive, spin-half particles from a spinning black hole has been studied with particular emphasis to the nature of the radial wave functions and the reflection and trans- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Scattering of Dirac Waves off Kerr Black Holes 9 Fig. 3a: Reflection (R) coefficient of waves with varying mass as functions of r∗. mp = 0.16 (solid), mp = 0.164 (dotted) and mp = 0.168 (long-dashed) are used. Other parameters are a = 0.95 and σ = 0.168. mission coefficients. Well known quantum mechanical step-potential approach is used but we applied it successfully to a complex problem of barrier penetration in a spacetime around a spinning black hole. Among significant observations, we find that the wave function and R, and T behave similarly close to the horizon independent of the initial parameter, such as the particle mass mp. Particles of different mass scatter off to a large distance completely differ- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 10 S. K. Chakrabarti and Banibrata Mukhopadhyay Fig. 3b: Amplitude of Re(Z+) of waves with varying mass as functions of r∗. mp = 0.16 (solid), mp = 0.164 (dotted) and mp = 0.168 (long-dashed) are used. Nature of potential with mp = 0.168 is drawn shifting vertically by 2.05 unit for clarity. Other parameters are a = 0.95 and σ = 0.168. ently, thus giving an impression that a black hole could be treated as a mass spectrograph! When the energy of the particle becomes higher compared to the rest mass, the reflection coefficient diminishes as it should it. Similar to a barrier penetration problem, the reflection coefficient becomes close to a hundred percent when the wave hits the potential barrier. Another significant observation is that the reflection and transmission coefficients are func- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Scattering of Dirac Waves off Kerr Black Holes 11 Fig. 4a: Contours of constant amplitude are plotted in the meridional plane around a black hole. Radial direction on equatorial plane is along X axis and the vertical direction and along Y . Both radial and theta solutions have been combined. Parameters are a = 0.5, mp = 0.8 and σ = 0.8. tions of the radial coordinates. This is understood easily because of the very nature of the potential barrier which is strongly space dependent which we approximate as a collection of steps. Combining with the solution of theta-equation, we find that the wave-amplitude vanishes close to the vertical axis, possibly due to the frame-dragging effects. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 12 S. K. Chakrabarti and Banibrata Mukhopadhyay Fig. 4b: Three dimensional view of R−1/2S−1/2 are plotted in the meridional plane around a black hole. Both radial and theta solutions have been combined. Parameters are a = 0.5, mp = 0.8 and σ = 0.8. REFERENCES Chandrasekhar 1976, Proc. R. Soc. Lond. A, 349, 571 Chakrabarti, S.K. 1984, Proc. R. Soc. Lond. A, 391, 27 Goldberg, J.N., Macfarlane, A.J., Newman, E.T., Rohrlich, F. & Sudarsan, E.C.G. 1967, J. Math. Phys., 8, 2155 Breure, R.A., Ryan, M.P. Jr. & Waller, S 1982, Proc. R. Soc. Lond. A, 358, 71 Chandrasekhar, S. 1983, in The Mathematical Theory Of Black Holes (London: Clarendon Press). Davydov, A.S. 1976 (Second Ed.) in Quantum Mechanics (Oxford, New York: Pergamon Press). c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
astro-ph/0702433
1
0702
2007-02-16T15:07:07
Nuclear Star Clusters (Nuclei) in Spiral Galaxies and Connection to Supermassive Black Holes
[ "astro-ph" ]
HST observations have revealed that compact sources exist at the centers of many, maybe even most, galaxies across the Hubble sequence. These sources are called "nuclei" or also "nuclear star clusters" (NCs), given that their structural properties and position in the fundamental plane are similar to those of globular clusters. Interest in NCs increased recently due to the independent and contemporaneous finding of three groups (Rossa et al. for spiral galaxies; Wehner & Harris for dE galaxies; and Cote et al. for elliptical galaxies) that NC masses obey similar scaling relationships with host galaxy properties as do supermassive black holes. Here we summarize the results of our group on NCs in spiral galaxies. We discuss the implications for our understanding of the formation and evolution of NCs and their possible connection to supermassive black holes.
astro-ph
astro-ph
Stellar Populations as Building Blocks of Galaxies Proceedings IAU Symposium No. 241, 2007 A. Vazdekis and R. Peletier., eds. c(cid:13) 2007 International Astronomical Union DOI: 00.0000/X000000000000000X Nuclear Star Clusters (Nuclei) in Spirals and Connection to Supermassive Black Holes Roeland P. van der Marel1, Joern Rossa2, Carl Jakob Walcher3, Torsten Boeker4, Luis C. Ho5, Hans-Walter Rix6, Joseph C. Shields7 1STScI, 3700 San Martin Drive, Baltimore, MD 21218, USA 2Dept. of Astronomy, University of Florida, Gainesville, FL 32611, USA 3Laboratoire d'Astrophysique de Marseille, F-13376 Marseille Cedex 12, France 4European Space Agency, Department RSSD, 2200 AG Noordwijk, The Netherlands 5Carnegie Institution of Washington, 813 Santa Barbara Street, Pasadena, CA 91101, USA 6Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany 7Dept. of Physics and Astronomy, Ohio University, Athens, OH 45701, USA Abstract. HST observations have revealed that compact sources exist at the centers of many, maybe even most, galaxies across the Hubble sequence. These sources are called "nuclei" or also "nuclear star clusters" (NCs), given that their structural properties and position in the fundamental plane are similar to those of globular clusters. Interest in NCs increased recently due to the independent and contemporaneous finding of three groups (Rossa et al. for spiral galaxies; Wehner & Harris for dE galaxies; and Cot´e et al. for elliptical galaxies) that NC masses obey similar scaling relationships with host galaxy properties as do supermassive black holes. Here we summarize the results of our group on NCs in spiral galaxies. We discuss the implications for our understanding of the formation and evolution of NCs and their possible connection to supermassive black holes. 1. Introduction High spatial resolution observations with the Hubble Space Telescope (HST) have re- vealed that many galaxies have a compact source in their very center. These "nuclei" have structural properties that are similar to those of globular clusters (see below), and they are therefore also called "nuclear star clusters" (NCs). Figure 1a shows the nearby Sd galaxy NGC 300 as an example. This galaxy has a NC, which is easily identified as a separate component because of the marked upturn in the surface brightness profile at small radii (Figure 1b). Boeker et al. (2002, 2004) observed a sample of 77 late-type spiral galaxies with HST/WFPC2 and found NCs in 77% of the sample. Fits to the two-dimensional images yield the effective radii and luminosities of the NCs, which for late-type spirals have median values reff = 3.5 pc and LI = 106.2L⊙. The study of NCs in late-type spirals is facilitated by the faintness of an underlying bulge/spheroid component. The same is true for the study of nuclei in dE galaxies. However, NCs are also found in other galaxy types. Carollo et al. (1997, 1998) report a detection frequency of 50 -- 60% in early-type spirals using HST/WFPC2 imaging. Cot´e et al. (2006) report the detection of nuclei in 66 -- 82% of intermediate-luminosity Virgo cluster elliptical galaxies (but no detections in the most luminous giant ellipticals) using HST/ACS imaging. The detection of NCs does become progressively more difficult as the density of the underlying spheroid increases, which is a problem especially for elliptical galaxies with steep cusps. The inferred NC characteristics then depend sensitively on the assumed spheroid brightness profile (Lauer et al. 2006; Ferrarese et al. 2006b). 1 7 0 0 2 b e F 6 1 1 v 3 3 4 2 0 7 0 / h p - o r t s a : v i X r a 2 van der Marel et al. Figure 1. (a) HST/ACS color composite image of the central 1.2 × 1.2 kpc of the Sd galaxy NGC 300 at D ≈ 1.8 Mpc (from the observations of Bresolin et al. 2005). A prominent nuclear star cluster (NC) resides at the center of the galaxy. (b) I-band surface brightness profile in mag/arcsec2 for the central ∼ 10′′ = 87 pc, as determined from HST/WFPC2 data of the same galaxy (Boeker et al. 2002). The NC is evident as a marked upturn in the brightness profile, as compared to inward extrapolations (solid, dashed lines) of the the brightness profile at large radii. King-model fits to the two-dimensional image imply a NC luminosity LI = 106.2L⊙ and effective radius reff = 2.9 pc (Boeker et al. 2004). (c) HST/STIS spectrum of the NC in NGC 300 (from Rossa et al. 2006). The red curve shows the best spectral population fit, which has Z = 0.004, AV = 0.4, a luminosity-weighted mean hlog(age/Gyr)i = 8.63 and M/LB = 0.51M⊙/LB,⊙. The implied NC mass is M = 105.7M⊙. 2. Spectroscopy of Nuclear Star Clusters in Spiral Galaxies NCs/nuclei are a common component of galaxies across the Hubble sequence. It is there- fore important to have observational knowledge of their stellar populations, ages, and masses. We pursued a spectroscopic program to measure these quantities for spiral galax- ies. We used the UVES echelle spectrograph on the VLT to obtain spectra of the NCs in nine late-type spirals (Walcher et al. 2005, 2006). The high spectral resolution of these data allows the measurement of the NC velocity dispersions from the near-IR Ca II triplet line broadening. The results fall in the range σ = 13 -- 34 km/s. When interpreted with spherically symmetric dynamical models tailored to fit the HST surface brightness profiles, these measurements yield the NC masses. Figure 2a shows the NC properties in a face-on fundamental plane projection. The NCs fall on a high-mass extension of the globular cluster sequence. The VLT ground-based spatial resolution contaminates the NC light with that of the underlying galaxy. This is manageable in the latest-type spirals, but makes it difficult to study NCs in earlier-type spirals. We therefore also pursued a program with HST/STIS of 40 spiral galaxies of various Hubble types (Rossa et al. 2006). These spectra have the advantage of higher spatial resolution that can cleanly separate the NC light from the underlying galaxy light. However, the S/N and spectral resolution are much lower than for the VLT spectra, which makes it impossible to infer the NC kinematics. We fitted the VLT and the STIS spectra using a non-negative weighted linear sum of Bruzual & Charlot (2003) single-age templates with different ages. This yields for each NC the extinction, metallicity, star formation history, luminosity-weighted mean log(age), and M/L. Figure 1c shows the STIS spectrum of NGC 300 to illustrate this approach. The M/L inferred from the spectral population fitting can be combined with the luminosity inferred from HST imaging to yield the NC mass. For the nine NCs with VLT spectra, we find that the M/L and M values inferred from dynamical modeling and spectral population fitting are in good agreement. The NCs have mixed-age populations, with single-age fits generally ruled out by χ2 statistics. This is supported by an analysis of line-strength indices measured from the VLT Nuclear Star Clusters in Spirals 3 Figure 2. (a) Mean projected mass density inside reff versus total mass, for different types of stellar systems as labeled in the plot (from Walcher et al. 2005). Elliptical galaxies and bulges are found on the right side of the plot, which presents an approximately face-on view of their fundamental plane. Globular clusters are found on the left, roughly along a locus of constant reff ≈ 3 pc (solid line). The nuclear star clusters (NCs) in spiral galaxies (solid squares) lie at the high-mass end of the globular cluster sequence, as do the nuclei of dE galaxies. (b) Luminosity-weighted mean log(age/Gyr) versus Hubble type. NCs in late-type spiral galaxies tend to have younger populations than those in early-type spiral galaxies. (c) NC mass versus B-band luminosity of the host galaxy bulge. Open symbols denote early-type spirals and solid symbols denote late-type spirals. Galaxies with more luminous bulges tend to have more massive NCs. Solid lines are least squares fits to the data points. Panels (b,c) are from Rossa et al. (2006), and are based on stellar population fits to HST/STIS spectra such as those shown in Figure 1c. The dashed line in the right panel indicates the relation between the supermassive black hole mass and bulge luminosity for galaxies studied by Marconi & Hunt (2003). spectra. Approximately half of the NCs in the STIS sample contain a population younger than 1 Gyr. The luminosity-weighted ages range from 10 Myrs to 10 Gyrs. The NCs in late-type spirals tend to have younger luminosity-weighted mean ages than those in early- type spirals (Figure 2b). But even when a young population dominates the light, we find that there is generally an underlying older population that contains most of the mass. Carollo et al. (2002) found that NC luminosities in early-type spirals are on average brighter than those in late-type spirals. Our results show that this is true for NC masses as well: they correlate loosely with total galaxy luminosity. Figure 2c shows that there is an even stronger correlation with the luminosity of the host galaxy bulge. This correlation has the same slope as the well-known correlation between supermassive black hole (BH) mass and bulge luminosity (dashed line in Figure 2c). 3. Discussion The results of the spectroscopic analysis provide insight into several important questions. What do we know about NC formation? NCs have similar structural properties as globular clusters. This suggests a commonality in the physical processes that shaped them, despite their different environments. By contrast, bulges follow very different scal- ing relationships, indicating that their formation was probably governed by different physical processes. NCs generally have underlying old stellar populations, suggesting that they may have formed early in the life of the galaxy. Moreover, NC formation must be efficient, because NCs/nuclei are a common feature of galaxies. However, there must be circumstances that prevent NC formation in some fraction of galaxies. How do NCs and the galaxies they reside in evolve? NCs have had repeated star formation episodes. So they grow in mass as time passes, due to supply of new gas to the galaxy center through secular evolution processes (e.g., bar torques). NCs in later-type galaxies on average have younger populations, probably due to their larger gas supply. By contrast, nuclei in dEs have predominantly old populations (Geha, Guhathakurta & 4 van der Marel et al. van der Marel 2003), probably reflecting their general lack of gas. In fact, dEs may have formed through transformation of late-type spirals (Moore, Lake & Katz 1998), so dE nuclei may be the more evolved counterparts of the NCs in spirals. NCs can also cause secular evolution, given that their masses can be sufficient to drive the destruction of bars and the formation of pseudo-bulges (Carollo 1999; Kormendy & Kennicut 2004). What is the relation between NCs and supermassive BHs? NC masses scale with host galaxy properties similarly as BH masses. Contemporaneously with our work on spirals (Rossa et al. 2006), this was independently shown also for the nuclei in dE galax- ies (Wehner & Harris 2006) and the nuclei in intermediate-luminosity elliptical galaxies (Cot´e et al. 2006). In all cases the NC masses are a few times larger than BH masses that one would predict based on established relations with Lbulge (see, e.g., Figure 2c) or σ. Ferrarese et al. (2006a) showed that better agreement is obtained when host galaxy mass is used as the independent variable. Also, BHs are typically measured in galaxies more massive than a few times 1010M⊙, whereas NCs are typically detected in less massive galaxies. This latter finding could indicate that they are different manifestations of a single "Compact Massive Object" class. But it could also be a selection effect; e.g., the dwarf Seyfert NGC 4395 has both a BH and a NC (Filippenko & Ho 2003). The impli- cations of all these findings are discussed in detail in each of the cited papers, including the interesting possibility that BHs and NCs are somehow causally or evolutionarily con- nected. However, this does not necessarily follow from the data. McLaughlin et al. (2006) discussed a feedback mechanism that can account for all observed relations without in- voking any direct connection between BHs and NCs. As of yet, robust insights into the true BH-NC connection are still mostly missing. Acknowledgements Support for HST proposals #9070 and #9783 was provided by NASA through a grant from STScI, which is operated by AURA, Inc., under NASA contract NAS 5-26555. References Boeker, T., Laine, S., van der Marel, R. P., Sarzi, M., Rix, H.W., et al. 2002, AJ, 123, 1389 Boeker, T., Sarzi, M., McLaughlin, D. E., van der Marel, R. P., et al. 2004, AJ, 127, 105 Bresolin, F., Pietrzynski, G., Gieren, W., & Kudritzki, R.-P. 2005, ApJ, 634, 1020 Bruzual, G., & Charlot, S. 2003, MNRAS, 344, 1000 Carollo, C. M., Stiavelli, M., de Zeeuw, P. T., & Mack, J. 1997, AJ, 114, 2366 Carollo, C. M., Stiavelli, M., & Mack, J. 1998, AJ, 116, 68 Carollo, C. M. 1999, ApJ, 523, 566 Carollo, C. M., Stiavelli, M., Seigar, M., de Zeeuw, P. T., & Dejonghe, H. 2002, AJ, 123, 159 Cot´e, P., et al. 2006, ApJS, 165, 57 Ferrarese, L., et al. 2006a, ApJ, 644, L21 Ferrarese, L., Cot´e, P., Blakeslee, J. P., Mei, S., Merritt, D., West, M. 2006b [astro-ph/0612139] Filippenko, A. V., & Ho, L. C. 2003, ApJ, 588, L13 Geha, M., Guhathakurta, P., & van der Marel, R. P. 2003, AJ, 126, 1794 Kormendy, J., & Kennicutt, R. C., Jr. 2004, ARA&A, 42, 603 Lauer, T., et al. 2006, ApJ, submitted [astro-ph/0609762] Marconi, A., & Hunt, L. K. 2003, ApJ, 589, L21 McLaughlin, D. E., King, A. R., Nayakshin, S. 2006, ApJ, 650, L37 Moore, B., Lake, G., Katz, N. 1998, ApJ, 495, 139 Rossa, J., van der Marel, R. P., Boeker, T., Gerssen, J., Ho, L. C., et al. 2006, AJ, 132, 1074 Walcher, C. J., van der Marel, R. P., McLaughlin, D. E., Rix, H.-W., et al. 2005, ApJ, 618, 237 Walcher, C. J., Boeker, T., Charlot, S., Ho, L. C., Rix, H, Rossa, J., et al. 2006, ApJ, 649, 692 Wehner, E.H., & Harris, W. 2006, ApJ, 644, L17
astro-ph/0104127
1
0104
2001-04-07T01:03:53
The Formation of Stellar Clusters: Mass Spectra from Turbulent Molecular Cloud Fragmentation
[ "astro-ph" ]
Turbulent fragmentation determines where and when protostellar cores form, and how they contract and grow in mass from the surrounding cloud material. This process is investigated, using numerical models of self-gravitating molecular cloud dynamics. Molecular cloud regions without turbulent driving sources, or where turbulence is driven on large scales, exhibit rapid and efficient star formation in a clustered mode, whereas interstellar turbulence that carries most energy on small scales results in isolated star formation with low efficiency. The clump mass spectrum of shock-generated density fluctuations in pure hydrodynamic, supersonic turbulence is not well fit by a power law, and it is too steep at the high-mass end to be in agreement with the observational data. When gravity is included in the turbulence models, local collapse occurs, and the spectrum extends towards larger masses as clumps merge together, a power-law description dN/dM ~ M^nu becomes possible with slope nu < -2. In the case of pure gravitational contraction, i.e. in regions without turbulent support, the clump mass spectrum is shallower with nu = -3/2. The mass spectrum of protostellar cores in regions without turbulent support and where turbulence is replenished on large-scales, however, is well described by a log-normal or by multiple power laws, similar to the stellar IMF at low and intermediate masses. In the case of small-scale turbulence, the core mass spectrum is too flat compared to the IMF for all masses.
astro-ph
astro-ph
(ACCEPTED FOR PUBLICATION IN APJ) Preprint typeset using LATEX style emulateapj v. 04/03/99 THE FORMATION OF STELLAR CLUSTERS: MASS SPECTRA FROM TURBULENT MOLECULAR CLOUD FRAGMENTATION UCO/Lick Observatory, University of California at Santa Cruz, Santa Cruz, CA 95064, USA (e-mail: [email protected]) Max-Planck-Institut für Astronomie, Königstuhl 17, 69117 Heidelberg, Germany RALF S. KLESSEN 1 0 0 2 r p A 7 1 v 7 2 1 4 0 1 0 / h p - o r t s a : v i X r a (Accepted for publication in ApJ) ABSTRACT Star formation is intimately linked to the dynamical evolution of molecular clouds. Turbulent fragmentation determines where and when protostellar cores form, and how they contract and grow in mass via competitive accre- tion from the surrounding cloud material. This process is investigated, using numerical models of self-gravitating molecular cloud dynamics, where no turbulent support is included, where turbulence is allowed to decay freely, and where it is continuously replenished on large, intermediate and small scales, respectively. Molecular cloud regions without turbulent driving sources, or where turbulence is driven on large scales, exhibit rapid and efficient star formation in a clustered mode, whereas interstellar turbulence that carries most energy on small scales results in isolated star formation with low efficiency. The clump mass spectrum of shock-generated density fluctuations in pure hydrodynamic, supersonic turbulence is not well fit by a power law, and it is too steep at the high-mass end to be in agreement with the observational data. When gravity is included in the turbulence models, local collapse occurs, and the spectrum extends towards larger masses as clumps merge together, a power-law description dN/dM ∝ Mν becomes possible with slope ν ∼< - 2. In the case of pure gravitational contraction, i.e. in regions without turbulent support, the clump mass spectrum is shallower with ν ≈ - 3/2. The mass spectrum of protostellar cores in regions without turbulent support and where turbulence is replen- ished on large-scales, however, is well described by a log-normal or by multiple power laws, similar to the stellar IMF at low and intermediate masses. The model clusters are not massive enough to allow for comparison with the high-mass part of the IMF. In the case of small-scale turbulence, the core mass spectrum is too flat compared to the IMF for all masses. Subject headings: hydrodynamics – ISM: clouds – ISM: kinematics and dynamics – ISM: clump mass spectrum – stars: formation – stars: IMF – turbulence 1. INTRODUCTION Understanding the processes that lead to the formation of stars is one of the fundamental challenges in astronomy. As mass is the dominant parameter determining stellar evolution, reproducing and explaining the initial mass function of stars (IMF) is a key requisite for any realistic theory of star forma- tion. Stars are born in turbulent interstellar clouds of molecular hydrogen. The location and the mass growth of young stars are hereby intimately coupled to the dynamical cloud environment. Stars form by gravitational collapse of shock-compressed den- sity fluctuations generated from the supersonic turbulence ubiq- uitously observed in molecular clouds (e.g. Elmegreen 1993, Padoan 1995, Klessen, Heitsch, & Mac Low 2000, Padoan et al. 2001). Once a gas clump becomes gravitationally unstable, it begins to collapse and the central density increases consider- ably, giving birth to a protostar. In this dynamic picture, star formation takes place roughly on a free-fall timescale, as op- posed to the "standard" model of the inside-out collapse of sin- gular isothermal spheres, where core formation is dominated by the ambipolar diffusion timescale (Shu 1977, Shu, Adams, & Lizano 1987). Altogether, star formation can be seen as a two-phase process: First, turbulent fragmentation leads to tran- sient clumpy molecular cloud structure, with some of the den- sity fluctuation exceeding the critical mass and density for grav- itational contraction. Second, the collapse of individual Jeans- unstable protostellar clumps builds up the stars. In this phase, a nascent protostar grows in mass via accretion from the infalling 1 envelope until the available gas reservoir is exhausted or stellar feedback effects become important and remove the parental co- coon - a new star is born (e.g. André, Ward-Thompson, & Bar- sony 2000, Myers, Evans, & Ohashi 2000). The terms shock- generated density fluctuations and gas clumps are used synony- mously, and clumps are identified using a 3-dimensional clump- finding algorithm comparable to the one described in Williams, De Geus, & Blitz (1994). Protostellar cores in the simulations are defined as the (unresolved) high-density central regions of collapsing clumps, where individual protostars build up. Stars form in small aggregates or larger clusters (Lada 1992, Mizuno et al. 1995, Testi, Palla, & Natta 1998, also Adams & Myers 2001), where the interaction of protostellar cores and their competition for mass from their surrounding may become important for shaping the distribution of the final star proper- ties (Bonnell et al. 1997, Klessen, Burkert, & Bate 1998, and Klessen & Burkert 2000, 2001, in the following papers I and II, respectively). This complex evolutionary sequence involves a wide variety of different physical phenomena, and it is not at all well understood which processes dominate and determine the stellar mass spectrum. The current investigation is the fourth in a series which fo- cuses on the first phase of the star formation process, modeling the turbulent fragmentation of large subvolumes inside molec- ular clouds and the dynamical evolution towards the formation of clusters of protostellar cores. Numerical simulations (§2) of self-gravitating isothermal gas, without turbulent support, with decaying turbulence, and with supersonic compressible turbu- lence that is driven on large, intermediate, and small scales, are 2 Ralf S. Klessen TABLE 1 PROPERTIES OF THE CONSIDERED MOLECULAR CLOUD MODELS model type 1 2 3 4 5 Gaussian density decaying turbulence driven turbulence driven turbulence driven turbulence ka - [1 . . .8] 1 . . .2 3 . . .4 7 . . .8 Mrms b - [5.0] 3.3 3.3 3.3 c NJ 220 220 120 120 120 particles 5 × 105 2 × 105 2 × 105 2 × 105 2 × 105 further referenced model I in KB - model A1h in KHM model A2h in KHM model A3h in KHM aDriving wavenumber (in model 2, driving stopped when gravity is 'turned on' at the stage of fully established turbulence) bMach number in turbulent equilibrium as calculated from the 1-dimensional rms velocity dispersion, Mrms ≡ σ1D/cs, where cs is the sound speed and σ1D is defined via σ2 3D = 1/3 (Ekin/2) with Ekin being the total kinetic energy. Recall that the total mass is unity. For model 2 the value corresponds to time t < 0 before the driving mechanism was 'turned off' and turbulence was allowed to decay. 1D = 1/3 σ2 cNumber of (spherical) mean thermal Jeans masses contained in the system; note that this number is lower by 2 when using a cubic definition of the Jeans mass dCorresponding model name in KB (paper I) and in KHM (Klessen et al. 2000) for further details used to analyze the star formation resulting from the interplay between gravity on the one side and gas pressure and turbulent motions on the other (§3). In particular, the relation between the masses of molecular clumps, protostellar cores and the final stars in the considered models are discussed (§4) and the results summarized (§5). 2. NUMERICAL METHOD AND DRIVEN TURBULENCE To adequately describe turbulent fragmentation and the for- mation of protostellar cores, it is necessary to resolve the col- lapse of shock-compressed regions over several orders of mag- nitude in density. Due to the stochastic nature of supersonic turbulence, it is not known in advance where and when local collapse occurs. Hence, SPH (smoothed particle hydrodynam- ics) is used to solve the equations of hydrodynamics. It is a La- grangian method, where the fluid is represented by an ensem- ble of particles and flow quantities are obtained by averaging over an appropriate subset of the SPH particles (Benz 1990). The method is able to resolve large density contrasts as parti- cles are free to move and so naturally the particle concentration increases in high-density regions. SPH can also be combined with the special-purpose hardware device GRAPE (Sugimoto et al. 1990, Ebisuzaki et al. 1993; also Steinmetz 1996) permit- ting calculations at supercomputer level on a normal worksta- tion. The simulations presented here concentrate on subregions within a much larger cloud, therefore periodic boundary condi- tions are adopted (Klessen 1997). Once the high-density, pro- tostellar cores in the centers of collapsing gas clumps exceed a density limit four orders of magnitude above the mean density, they are substituted by 'sink' particles (Bate, Bonnell, & Price 1995). These particles have the ability to accrete gas from their envelopes, while keeping track of mass and linear and angular momentum. By adequately replacing high-density core with sink particles one is able to follow the dynamical evolution of the system over many free-fall times. The large observed linewidths in molecular clouds imply the presence of supersonic velocity fields that carry enough energy to counterbalance gravity on global scales (Williams, Blitz, & McKee 2000). However, it is known that turbulent energy dissi- pates rapidly, i.e. roughly on the free-fall timescale (Mac Low et al. 1998, Stone, Ostriker, & Gammie 1998, Padoan & Nordlund 1999). To prevent or considerably postpone global collapse, turbulence is required to be continuously replenished. This is achieved here by applying a non-local driving scheme, that in- serts energy in a limited range of wavenumbers such that the total kinetic energy contained in the system remains constant and compensates the gravitational contraction on global scales (Mac Low 1999). The models do not include magnetic fields, as their presence cannot halt the decay of turbulence (Mac Low et al. 1998, Stone et al. 1998, Padoan & Nordlund 1999) and does not significantly alter the efficiency of local collapse for driven turbulence (Heitsch, Mac Low, & Klessen 2001). Furthermore, possible feedback effects from the star formation process itself (like bipolar outflows, stellar winds, or ionizing radiation from new-born O or B stars) are neglected. This necessarily limits the interpretation of the mass spectra at very late evolutionary stages of the system. Hence, the current analysis restrains itself to phases when the mass accumulated protostellar cores is less than ∼ 70% of the total mass in the considered volume. Altogether, five different models of molecular cloud dynam- ics are considered here: To compare with the case of driven turbulence, model 1 describes the dynamical evolution of an initially Gaussian density fluctuation field where turbulence is assumed to have already decayed in the considered molecular cloud region. This describes the most extreme case of clustered star formation, and the simulation is identical to model I in pa- per I. The power spectrum of the initial density fluctuations is P(k) ∝ k- 2. Model 2 starts with a fully established supersoni- cally turbulent velocity field, but turbulence is allowed to decay freely. Driven turbulence is represented by model 3 where the energy source acts at wavenumbers k in the range 1 ≤ k ≤ 2, by model 4 which has 3 ≤ k ≤ 4, and model 5 with 7 ≤ k ≤ 8. The wavelengths of the corresponding perturbations are ℓ = L/k, where L is the total size of the computed volume. Hence, ki- Mass Spectra from Turbulent Fragmentation 3 In the adopted scaling, netic energy is continuously added on large, intermediate, and small scales, respectively. The driving strength is adjusted to yield the same constant turbulent Mach number Mrms = 3.3 for all three models (see Klessen et al. 2000, models A1h, A2h, and A3h). Turbulence that is driven on large scales appears to yield most appropriate description of molecular cloud dynam- ics and star formation as is suggested by statistical analysis of molecular cloud structure (e.g. Ossenkopf & Mac Low 2001). The relevant model parameters are listed in table 1. The models presented here are computed in normalized units. If scaled to mean densities of n(H2) = 105 cm- 3, a value typical for star-forming molecular cloud regions (e.g. in ρ-Ophiuchus, see Motte, André, & Neri 1998) and a temperature of 11.4 K (i.e. a sound speed cs = 0.2 km s- 1), the total mass contained in the computed volume in models 1 and 2 is 220 M⊙ and the size of the cube is 0.34 pc. This corresponds to 220 ther- mal Jeans masses. Models 3 to 5 have a mass of 120 M⊙ within a volume of (0.29 pc)3, equivalent to 120 thermal Jeans masses1. the mean thermal Jeans mass in all models is thus hMJi = 1 M⊙, the global free-fall timescale is τff = 105 yr, and the simulations cover a density range from n(H2) ≈ 100 cm- 3 in the lowest density regions to n(H2) ≈ 109 cm- 3 where collapsing protostellar cores are iden- tified and converted into 'sink' particles in the code. In this den- sity regime gas cools very efficiently and it is possible to use an effective polytropic equation-of-state in the simulations instead of solving the detailed radiation transfer equations. The effec- tive polytropic index is typically close to unity, γeff ∼< 1, except for densities 105 cm- 3 < n(H2) < 107 cm- 3, where smaller val- ues of γeff are expected (Spaans & Silk 2000). For simplicity, a value of γeff = 1, i.e. an isothermal equation of state, is adopted for all densities in the simulations. Concerning the gas tem- perature, this approximation is certainly valid as in star form- ing clouds the temperature cannot drop significantly below the adopted canonical value of 10 K, even in the regime 105 cm- 3 < n(H2) < 107 cm- 3. However, it needs to be noted that the stiff- ness of the equation of state also determines the density contrast in shock compressed gas, and hence influences the overall den- sity distribution in supersonic flows. For further discussions see Scalo et al. (1998), Ballesteros-Paredes, Vázquez-Semadeni, & Scalo (1999), and Spaans & Silk (2000). Variations in γeff influ- ence the local Jeans scale in shock-compressed density fluctua- tions and may modify the resulting mass spectrum of collapsing cores. This effect needs to be investigated in more detail. 3. STAR FORMATION FROM TURBULENT FRAGMENTATION Stars form from turbulent fragmentation of molecular cloud material. Supersonic turbulence, even if it strong enough to counterbalance gravity on global scales, will usually provoke local collapse. Turbulence establishes a complex network of interacting shocks, where converging shock fronts generate clumps of high density. This density enhancement can be large enough for the fluctuations to become gravitationally unstable and collapse. This happens when the local Jeans length be- comes smaller than the size of the fluctuation. However, the fluctuations in turbulent velocity fields are highly transient. The random flow that creates local density enhancements can dis- perse them again. For local collapse to actually result in the for- mation of stars, Jeans-unstable shock-generated density fluctua- tions must collapse to sufficiently high densities on time scales shorter than the typical time interval between two successive shock passages. Only then are they able to 'decouple' from the ambient flow and survive subsequent shock interactions. The shorter the time between shock passages, the less likely these fluctuations are to survive. Hence, the timescale and ef- ficiency of protostellar core formation depend strongly on the wavelength and strength of the driving source (Klessen et al. 2000, Heitsch et al. 2001), and accretion histories of individual protostars are strongly time varying (Klessen 2001, hereafter paper III). The velocity field of long-wavelength turbulence is found to be dominated by large-scale shocks which are very efficient in sweeping up molecular cloud material, thus creating massive coherent structures. When a coherent region reaches the criti- cal density for gravitational collapse its mass typically exceeds the local Jeans limit by far. Inside the shock compressed region, the velocity dispersion is much smaller than in the ambient tur- bulent flow and the situation is similar to localized turbulent decay. Quickly a cluster of protostellar cores forms. Therefore, models 1 to 3 with zero support, decaying and large-scale turbu- lence, respectively, lead to a clustered mode of star formation. The efficiency of turbulent fragmentation is reduced if the driv- ing wavelength decreases. When energy is inserted mainly on small spatial scales, the network of interacting shocks is very tightly knit, and protostellar cores form independently of each other at random locations throughout the cloud and at random times. Individual shock generated clumps have lower mass and the time interval between two shock passages through the same point in space is small. Hence, collapsing cores are easily de- stroyed again and star formation is inefficient. This scenario corresponds to the isolated mode of star formation. This is visualized in figure 1, showing the density structure of all five models at t = 0 and at a time when the first protostel- lar cores have formed by turbulent fragmentation and have ac- creted roughly 30% of the total mass. For model 1 (Gaussian), without turbulent support, the figure indicates at t = 0 the initial density distribution. For the other models, t = 0 corresponds to the phase of fully developed turbulence just before gravity is 'switched on' (in model 2 the driving mechanism is 'switched off' at the same time). Time is measured in units of the global free-fall timescale τff = (3π/32G)- 1/2 hρi - 1/2, with hρi being the mean density and G the gravitational constant. Dark dots indi- cate the location of dense collapsed core. In the non-supported model 1 all spatial modes are unstable initially and in model 2 of decaying turbulence they become unstable after roughly one crossing time. Therefore, these systems evolve into a fil- amentary structure and protostellar cores form predominantly at the intersections of the filaments. Similarly, also large-scale turbulence builds up a network of filaments, however, this time the large coherent structures are not caused by gravity, but in- stead are due to shock compression. Once gravity is included, it quickly dominates the evolution inside the shock compressed regions. The random velocity component is quickly damped by dissipation, and again a cluster of protostellar cores builds up. In the case of intermediate-wavelength turbulence, cores form in small aggregates, and small-scale turbulence predominantly results in the formation of isolated cores. Note the different times needed for mass to be accumulated in dense cores. 1Throughout this paper the spherical definition of the Jeans mass is used, MJ ≡ 4/3 πρλ3 , where G and R are the gravitational and the gas constant. The mean Jeans mass hMJi is then determined from average density in the system hρi. An alternative cubic definition, MJ ≡ ρ(2λJ)3, would yield a value roughly twice as large. J , with density ρ and Jeans length λJ ≡ (cid:0) πRT Gρ (cid:1)1/2 4 Ralf S. Klessen DE 1 DE 2  = 0:0 Ga ia  = 0:0 de ayig  = 0:0  = 0:0    = 1:1 Ga ia  = 1:3 de ayig  = 28:6  = 29:3   DE 3 DE 4 DE 5  = 0:0 k = 1 : : : 2  = 0:0 k = 3 : : : 4  = 0:0 k = 7 : : : 8  = 0:0  = 0:0  = 0:0     = 1:3 k = 1 : : : 2  = 1:8 k = 3 : : : 4  = 4:0 k = 7 : : : 8  = 26:7  = 31:5  = 28:6    FIG. 1.- Comparison of the gas distribution in the five models at two different phases of the dynamical evolution, at t = 0 indicating the initial density structure, just before gravity is 'switched on', and after the first cores have formed and accumulated roughly M∗ ≈ 30% of the total mass. The high-density (protostellar) cores are indicated by black dots. Note the different time interval needed to reach the same dynamical stage. Time is normalized to the global free-fall timescale of the system, which is τff = 105 yr for T = 11.4 K and n(H2) = 105 cm- 3. The cubes contain masses of 220 hMJi (models 1 and 2) and 120 hMJi (models 3 to 5), respectively, where the average thermal Jeans mass is hMJi = 1 M⊙ with the above scaling. The considered volumes are (0.32 pc)3 and (0.29 pc)3, respectively. Note, however, that the isothermal models are freely scalable as discussed in §2. Mass Spectra from Turbulent Fragmentation 5 4. MASS SPECTRA FROM TURBULENT FRAGMENTATION 4.1. The Mass Spectra Mass is the dominant parameter that determines stellar evolu- tion. It is therefore important to investigate the relation between the mass of molecular clumps and protostellar cores, and the stars resulting from the collapse of the former. For the five mod- els considered here, figure 2 plots the mass distribution of gas clumps (thin lines), of the subset of Jeans-critical clumps (thin lines, hetched distribution), and of collapsed cores (thick lines, hetched area). It depicts four different evolutionary phases, the initial distribution just when gravity is 'switched on' (at t = 0, left column), and then after (turbulent) fragmentation has led to protostellar core formation, i.e. when the fraction M∗ of the total mass accumulated in dense cores has reached values of M∗ ≈ 5%, M∗ ≈ 30%, and M∗ ≈ 60% (columns 2 to 4, respec- tively). The clump mass spectra are obtained applying a clump- finding algorithm similar to the one described by Williams et al. (1994), but working on all three spatial coordinates and adapted to make use of the SPH kernel smoothing procedure (for details see Appendix 1 in paper I). To guide your eye, two dotted lines indicate a slope ν = - 1.5 typical for the observed power-law clump mass spectrum dN/dM = Mν, as well as the Salpeter (1955) approximation ν = - 2.33 to the stellar IMF appropri- ate for intermediate and high masses (e.g. Scalo 1998, Kroupa 2001). 4.2. Time Evolution of Clump Mass Spectra in the Interplay between Turbulence and Self-Gravity In the initial phase, i.e. before gravity is 'turned on' and local collapse begins to set in, the clump mass spectrum (thin line) is not well described by a single power law. The distribution has small width and falls off steeply at larger masses. Below masses M ≈ 0.3 hMJi the distribution becomes shallow, and strongly declines at and beyond the resolution limit (indicated by a ver- tical line). Clumps are on average considerably smaller than the mean Jeans mass in the system hMJi. For masses M > 0.1 hMJi and for models 1 to 4, this behavior resembles the spectrum of pre-stellar condensations found in ρ-Ophiuchus (Motte et al. 1998, Johnstone et al. 2000, see also Testi & Sargent 1998 for Serpens). Recall that for densities of n(H2) = 105 cm- 3 and temperatures T = 11.4 K, the mean Jeans mass in the system is hMJi = 1 M⊙. In the later evolution the effects of gravitational attraction modify the distribution of clump masses. Clumps merge and grow bigger, and the mass spectrum extends towards larger masses. At the same time the number of cores which exceed the Jeans limit increases. Local collapse sets in and results in the formation of dense cores. This happens fastest and is most evident in model 1 which lacks turbulent support. The velocity field is entirely determined by gravitational contraction on all scales and at all times. The clump mass spectrum is very well fit by a single power law and exhibits a slope ν ≈ - 1.5 as long as protostellar cores are forming and the overall gravitational potential is dominated by non-accreted gas. The influence of gravity on the clump mass distribution is weaker where turbulence dominates over gravitational contrac- tion on the global scales, i.e. in model 2 during the early stages and in models 3 to 5 during all phases. The more the turbulent energy dominates over gravity, the more the spectrum resem- bles the initial case of pure hydrodynamic turbulence. This is most extreme in model 5 of small-wavelength turbulence, where the short interval between shock passages prohibits ef- ficient clump merging and the build up of a large number of massive clumps. Only few clumps exceed the Jeans limit, be- come gravitationally unstable, and collapse to form cores. The bulk of the mass distribution remains unchanged by gravity and is never well fit by a single power law. The mass spectrum re- tains the initial shape with only few collapsed clumps added at the high-mass end. When the scalelength of the dominant turbulent mode is increased, the density structure becomes more coherent and clump mergers are more frequent. The number of high-mass Jeans-unstable clumps increases, yielding a wider clump mass distribution which exhibits a power-law behavior for all masses larger than the resolution limit. In models 2 to 4, the slope lies in the interval - 2.5 < ν ∼< - 2. For individual models in fig- ure 2, the slope ν increases with time as the statistical proper- ties of the system become more and more influenced by clump merging and gravitational contraction onto high-density cores. When comparing similar evolutionary phases for different mod- els, again the clump spectrum falls off less steeply if gravity dominates the evolution over larger spatial scales, i.e. ν de- creases from model 1 to model 5. When M∗ ≈ 60% ν ≈ - 1.5 for model 1, ν ∼< - 2 for models 2 and 3, while for models 4 and 5 a power-law description is no longer sensible. In sum- mary, the clump mass spectrum gets shallower when gravity be- comes more important. This could explain the observed range of slopes - 1.9 ∼< ν ∼< - 1.3 for the clumps mass spectra in dif- ferent molecular cloud regions (e.g. Stutzki & Güsten 1990, Williams et al. 1994, Heithausen et al. 1998, Kramer et al. 1998, Onishi et al. 1998). The importance of self-gravity for shap- ing the velocity and density structure may differ from cloud to cloud. In the case of strong gravity, their statistical proper- ties furthermore depend sensitively on the viewing angle (e.g. Klessen 2000). In the late evolutionary stages, similar behavior holds for the subset of Jeans-unstable clumps. For the three models of driven turbulence, the distribution of the gravitationally super- critical clumps (as indicated by hatched thin lines) is largest and widest for model 3 and decreases in width and size towards model 5. The clump spectra of models 2 and 3 are similar, indicating that the conditions for local collapse within the co- herent shock-compressed regions which result from large-scale driving turbulence are comparable to turbulent decay. Within these coherent structures, turbulence is strongly reduced and stars formation is efficient. In the early stages of the evolu- tion, however, the mass spectrum of Jeans-unstable clumps is not well described by a power law for all models, instead it is more compatible with a log-normal distribution. For the Gaus- sian model the peak is roughly at the average thermal Jeans mass hMJi and decreases towards smaller masses when includ- ing turbulence and decreasing the driving wavelength. The dis- tribution of Jeans-unstable clumps in model 5 peaks roughly at 1/4 hMJi. Thus small-scale turbulence produces clumps of on average smaller mass scale than does large-scale turbulence. 4.3. Protostellar Mass Spectra from Turbulent Fragmentation Like the distribution of Jeans-unstable clumps, also the mass spectrum of dense protostellar cores (thick hatched line) resem- bles a log-normal in the model without turbulent support, and in the ones with turbulent decay or long-wavelength turbulent driving. A log-normal fit is obtained at times M∗ ≈ 30% and M∗ ≈ 60%, and indicated by long-dashed lines in column 3 and 4. The corresponding mean values and widths are given in table 6 Ralf S. Klessen FIG. 2.- Mass spectra of gas clumps (thin lines), and of the subset of Jeans unstable clumps (thin lines, hatched distribution), and of dense collapsed cores (hatched thick-lined histograms). Masses are binned logarithmically and normalized to the average thermal Jeans mass hMJi. The left column gives the initial state of the system, just when gravity is 'switched on', the second column shows the mass spectra when M∗ ≈ 5% of the mass is accreted onto dense cores, the third column describes M∗ ≈ 30%, and the last one M∗ ≈ 30%. For comparison with power-law spectra (dN/dM ∝ Mν ), a slope ν = - 1.5 typical for the observed clump mass distribution, and the Salpeter slope ν = - 2.33 for the IMF at intermediate and large masses, are indicated by the dotted lines in each plot. Note that with the adopted logarithmic mass binning these slopes appear shallower by +1 in the plot. The vertical line shows the SPH resolution limit. In columns 3 and 4, the long dashed curve shows the best log-normal fit to the core mass spectrum. To compare the distribution of core masses with the stellar IMF, an efficiency factor of roughly 1/3 to 1/2 for the conversion of protostellar core material into single stars needs to be taken into account, as discussed in the text. For T = 11.4 K and n(H2) = 105 cm- 3, the average Jeans mass in the system is hMJi = 1 M⊙. Note, however, that the considered models can be scaled to different hMJi as well. Mass Spectra from Turbulent Fragmentation 7 2. As for the Jeans-critical cores, the peak is roughly at the av- erage thermal Jeans mass hMJi of the system. The width of the distribution spans two orders of magnitude for the cluster size considered here, and is approximately the same for all five mod- els. However, a log-normal fit is only appropriate for models 1 to 3. The core mass spectrum for models 4 and 5 is too flat, and a fit is not attempted. Here the accretion histories of individual protostellar cores are not well correlated, i.e. gas clumps typi- cally do not contain multiple protostellar cores. In this isolated mode of star formation, mutual interaction and competition for gas accretion are not important. A log-normal shape of the mass distribution may be ex- plained by invoking the central limit theorem (e.g. Larson, 1973, Zinnecker 1984, also Adams & Fatuzzo 1996), as pro- tostellar cores form and evolve through a sequence of highly stochastic events, resulting from the statistical nature of super- sonic turbulence, stochastic clump merging, and/or competitive accretion within merged clumps containing multiple cores. The fact that the quality of the log-normal description becomes bad for the case of isolated star formation (model 5) indicates that clump merging and competitive accretion are dominant factors leading to a log-normal mass spectrum. Isolated star formation in supersonic turbulence exhibits a featureless flat mass spec- trum. It appears that the collective effects of gravitational col- lapse and the synchronization of individual accretion histories (i.e. the mutual interaction of protostellar cores and their com- petition for accretion from a common gas reservoir) are nec- essary to obtain mass spectra with features similar to the IMF (§4.4). The fact that the distribution peaks roughly at the aver- age Jeans mass, despite strong variations of the local Jeans mass for individual clumps which span a wide range of densities, in- dicates that the system retains in a statistical sense knowledge of its mean properties, even in the case of supersonic turbu- lence. The same behavior is seen in paper's I and II for a wide varyity of Gaussian initial conditions. It is the average thermal Jeans mass that introduces a scale into the mass function (e.g. Larson 1998, Elmegreen 1999). 4.4. Comparison with the Stellar Mass Function For the direct comparison between the core mass spectrum and the IMF one needs to adopt a physical scaling for the nu- merical model (§2). This is done here by taking the average thermal Jeans mass to be hMJi = 1 M⊙. Furthermore, one needs to estimate which fraction of a typical protostellar core in the model will accrete onto the star(s) in its interior, and which frac- tion may be expelled during the main accretion phase by pro- tostellar outflows, or be removed by tidal effects in a clustered environment. For isolated cores forming single stars, the mass loss due to radiation or outflows is expected to be small and most of the core material will indeed end up in stars (Wuchterl & Tscharnuter 2000). In this case the core mass spectrum can be compared directly to the single-star IMF. If cores con- tain a large amount of angular momentum, they are likely to form binary stars. Indeed a initial binary fraction of almost 100% is consistent with observations of star clusters (Kroupa 1995). Assuming a more or less uniform distributon of mass ratios (Duquennoy & Mayor 1991) leads to a shift of a factor of two in the characteristic mass of the distribution compared to the single-star IMF. If the number of triple systems is high, this shift could be even larger. To some degree this effect can be taken into account by comparing the core mass spectrum with IMF estimates that do not contain binary corrections (see Kroupa 2001 for a further discussion). In addition, a fraction of the accreting matter may settle into a protobinary disk and may not accrete onto the stars due to angular momentum conserva- tion (e.g. Bate 2000). In a cluster environment, this disk may be truncated and leak out matter due to tidal interactions (e.g. Clarke, Bonnell, & Hillenbrand 2000). These effects depend strongly on the stellar density of the cluster and its dynami- cal evolution. Altogether, the peak in the mass spectrum of protostellar cores is expected to exceed the characteristic mass in the single star IMF by a factor of 2 to 3 (i.e. is shifted by 0.3 ∼< ∆log10 M ∼< 0.5). Additional uncertainty stems from the possible formation of O or B stars in the stellar cluster. These would trigger the com- plete gas removal due to ionizing radiation, therefore limiting the core formation efficiency M∗. As it is not known in ad- vance whether at all and when high mass stars form during the protostellar cluster evolution, the current models should be con- sidered with caution at very late phases M∗ > 70%. If one assumes a close correspondence between core masses and stellar masses, then for models 1 to 3 the core mass dis- tribution compares well with the observed IMF at low to inter- mediate masses. The width of the distribution falls right in be- tween the log-normal IMF estimates by Miller & Scalo (1979) on one side and by Kroupa et al. (1990) and Scalo (1986) on the other. The same holds when using the more common mul- tiple power-law description of the IMF (Scalo 1998 or Kroupa 2001, see e.g. his figure 14). Also the characteristic masses in the distribution become comparable if one adopts values of few × 0.1 for the accretion efficiency from cores to the central stars taking into account unresolved binaries and possible tidal truncation of protobinary disks in a dense cluster environment as mentioned before. However, more work needs to be done to obtain better estimates for this efficiency factor. This requires combining the simulations described here with detailed dynam- ical models of pre-main sequence evolution, and extending the current scheme to resolving larger density contrasts and larger clusters sizes to be able to study in more detail the influence of cluster environment on the dynamical evolution of accretion disks. The protostellar clusters discussed here contain between 50 and 100 cores. This allows for comparison with the IMF only around the characteristic mass scale, i.e. at low to intermedi- ate masses. The numbers are too small to study the very low- and high-mass end of the distribution. The same holds for the mass spectra discussed in paper I and II. For high-mass stars (M ∼> 5 M⊙) the log-normal and the multiple power-law de- scriptions of the IMF begin to differ significantly. The log- normal models predict too few high-mass stars, and the same may also be true at the very low-mass end of the IMF, in the brown dwarf regime (especially when taking binary corrections into account). Because of insufficient statistics at the extreme ends of the distribution, the current set of simulations cannot be used to distinguish between log-normal and power-law IMF models. The log-normal fit at low to intermediate mass in fig- ure 2 is therefore mainly attempted for the sake of simplicity, because only two parameters, the peak value and the width, are sufficient to characterize the distribution. Both can be conve- niently compared with the corresponding fit parameters for the stellar mass function in the considered mass range. The comparison reveals a striking contrast between the mod- els of turbulent decay or large-scale driving on the one side, and models of short- to intermediate-wavelength turbulence on the other. Models 1 to 3 lead to protostellar mass spectra that agree well with the observations, whereas the mass spectra derived 8 Ralf S. Klessen TABLE 2 PROPERTIES OF LOG-NORMAL FIT Parameter log10 µ (peak) log10 σ (width) model 1 ∼ 30% ∼ 60% 0.13 0.52 0.26 0.47 model 2 ∼ 30% ∼ 60% - 0.10 0.36 0.21 0.47 model 3 ∼ 30% ∼ 60% - 0.34 - 0.03 0.36 0.40 model 4 ∼ 30% ∼ 60% model 5 ∼ 30% ∼ 60% - - - - - - - - NOTE.- The log-normal fits are obtained for models 1 to 3 only for two different evolutionary stages, when M∗ ≈ 30% and M∗ ≈ 60%. The applied functional form is dN/d log10 M ∝ exp(cid:2)- 0.5(log10 M - log10 µ)2/(log10 σ)2(cid:3), with mass M, mean µ and width σ scaled to the average Jeans mass in the system hMJi. No fit has been obtained for models 4 and 5, the core mass spectrum is too flat and featureless. When scaling the current models to physical units, the width of the core mass distributon lies between the IMF estimates by Miller & Scalo (1979) and by Scalo (1986) and Kroupa et al. (1990), who derive log10 σ = 0.67 (their estimate with constant star formation rate over 12 × 109 years) and log10 σ = 0.38 (their model MS), respectively. Recall that hMJi = 1 M⊙ for n(H2) = 105 cm- 3 and T = 11.4 K. As the number of protostars in the simulated cluster is limited to 50 – 100, the comparison with the stellar IMF applies to low to intermediate-mass stars only. The statistics is not good enough for an investigation of the very low-mass and the high-mass end of the IMF where the log-normal parametrization fails. from models 4 and 5 compare only very poorly with the stellar IMF. They are too flat or equivalently too wide. As small- to intermediate-scale turbulence describe an isolated mode of star formation, this finding is consistent with the hypothesis that most stars form in aggregates or clusters (e.g. Adams & My- ers 2001). To further constrain the numerical models discussed here, it will be necessary to compute the dynamical evolution of star-forming regions with 1000 protostellar cores or more. Be- sides the width and characteristic mass of the core distribution, also the detailed slope at very low and very high masses and the apparent symmetry around the peak can then be included into the analysis. Finally, it needs to be noted that the current findings raise doubts about attempts to explain the stellar IMF from the turbulence-induced clump mass spectrum only (e.g. Elmegreen 1993, Padoan 1995, Padoan et al. 2001, Padoan & Nordlund 2001). Quite typically for star forming turbulence, the collapse timescale of shock-compressed gas clumps often is comparable to their lifetime (molecular cloud clumps appear to be very tran- sient, e.g. Bergin et al. 1997). This not only has important con- sequences for the overall star formation efficiency in turbulent clouds (Klessen et al. 2000), but more so for the collapse behav- ior of individual Jeans-unstable shock-generated gas clumps. While collapsing to form or feed protostars, clumps may loose or gain matter from interaction with the ambient turbulent flow. In a dense cluster environment, collapsing clumps may merge to form larger clumps containing multiple protostellar cores, which subsequently compete with each other for accretion form the common gas environment (Bonnell 1997, paper's I and II). The resulting distribution of clump masses in star forming re- gions strongly evolves in time (figure 2). In dense clusters, fur- thermore, close encounters between accreting protostars may become important leading to the expulsion of protostars from the gas rich environment (as illustrated in figure 11 of paper I). This terminates mass growth and if occuring frequently enough modifies the resulting IMF. The mass accretion rates onto indi- vidual protostars are highly stochastic and strongly depend on the cluster environment (paper III). For all this reasons, it is not possible to infer a one-to-one relation between the clump masses resulting from turbulent molecular cloud fragmentation and the stellar IMF. Given our limited understanding of in- terstellar turbulence and protostellar mass growth processes in dense clusters, the current investigation (which attempts to in- clude some of the above processes) leads to the conclusion that – although tempting in some cases – it is not appropriate to take a snapshot of the turbulent clump mass spectrum as describing the IMF. 5. SUMMARY Stars form from turbulent fragmentation of molecular cloud material. It is the relation between turbulent fragmentation, (lo- calized) gravitational collapse and star formation which is the focus of this paper. As mass is the most important stellar param- eter, particular interest lies in the mass spectra of gas clumps, of the subset of gravitationally unstable gas clumps, and of proto- stellar cores, the latter one being the direct progenitors of stars. For this purpose five numerical models of the evolution of self- gravitating isothermal molecular gas have been analyzed, span- ning the parameter range relevant for molecular cloud dynam- ics. In model 1 turbulent support is not included and gravity is the dominant force shaping the velocity and density struc- ture. In model 2, initially supersonic compressible turbulence is allowed to decay freely, and in models 3 to 5 supersonic tur- bulence is continuously replenished on large, intermediate and small scales, respectively, such that gravitational attraction is compensated on global scales. In these models gravity is con- sidered only after turbulent equilibrium is established. It has been shown in a previous study (papers I and II) that molecular cloud regions without turbulent support form dense clusters of stars, regardless of the initial density struc- ture, within roughly one global free-fall timescale. When grav- itational contraction has sufficient time to act, the clump mass spectrum is well approximated by a power law dN/dM ∝ Mν. The mass distribution of protostellar cores, however, is better described by a log-normal with properties similar to the observed IMF of multiple stellar systems for low and intermediate-mass stars. This analysis is extended here by more realistically consid- Mass Spectra from Turbulent Fragmentation 9 ering molecular cloud regions where turbulence is allowed to decay and where it is continuously driven. Decaying turbu- lence leads to clustered star formation much like in the case of pure gravitational contraction. Supersonic turbulence, even if it is strong enough to compensate gravity on large scales, will provoke local collapse in shock compressed regions. As effi- ciency and timescale of star formation depend sensitively on the strength and the spatial scale of energy input into the sys- tem, large-scale turbulence leads to clustered star formation on short timescales, whereas for small-scale turbulence stars form in isolation and with low efficiency. This is reflected in the overall clump mass spectrum. The clump mass spectrum of pure hydrodynamic turbulence is not well described by a single power law, its width is too small com- pared to the observed data and it is too steep at the high-mass end. This changes, when gravity is taken into account. Clumps merging and accumulation of matter through local collapse, lead to a clump mass spectrum that extends to larger masses and that exhibits power-law behavior. The more strongly the overall dynamical evolution is influenced by gravity and synchonized, coherent collapse behavior, the flatter the power spectrum be- comes. In the extreme case of pure gravitational contraction the clump mass distribution exhibits a slope ν ≈ - 1.5. For the case of turbulence decay and large-scale injected turbulence the slope is ν ∼< - 2 during the intermediate phases of the dynamical evolution. In the case of small-scale turbulence, the infuence of gravity is weak and the clump mass distribution remains steep, close to the spectrum of purely hydrodynamic turbulence (i.e. before gravity was 'switched on' in the model). The depen- dence of the slope of the clump mass spectrum on the rela- tive importance of gravity may explain the range of observed power-law indices in different molecular clouds regions, as one expects the ratio between self-gravity and turbulent kinetic en- ergy to vary from cloud to cloud. Molecular cloud properties that result in clustered star forma- tion lead a stellar mass spectrum that is well fit by a log-normal at low and intermediate masses. The distribution exhibits a maximum close to the average Jeans mass in the system. This indicates that the system somehow retains 'knowledge' of its mean properties, even in the case of supersonic compressible turbulence. The mean thermal Jeans mass in a cloud indeed introduces a characteristic mass scale to clustered star forma- tion. For regions where turbulence is decaying or driven on large scales only, it appears that the collective effects of gravi- tational collapse and the correlation between individual accre- tion histories are necessary to obtain mass spectra with features similar to the IMF. Isolated star formation, on the contrary, as implied by turbulence that is driven on small scales, yields a featureless flat spectrum. This is in agreement with the hypoth- esis, that most stars form in aggregates and clusters. Shock-generated clumps in interstellar turbulence are highly transient. Their average lifetime in the turbulent flow is of the same order of their collapse timescale. In a dense cluster envi- ronment, furthermore competitive accretion and mutual proto- stellar interactions are important effects. The current investiga- tion shows that it is therefore not possible to infer a one-to-one relation between turbulent clump mass spectra and the stellar IMF. I thank Peter Bodenheimer, Andreas Burkert, Fabian Heitsch, Pavel Kroupa, Doug Lin, Mordecai-Mark Mac Low, and Günther Wuchterl for many stimulating discussions on star formation and the IMF and/or fruitful collaboration. I fur- thermore appreciate the comments and suggestions of the ref- eree John Scalo. They helped to clarify the arguments pre- sented here. I acknowledge financial support by a Otto-Hahn- Stipendium from the Max-Planck-Gesellschaft and partial sup- port through a NASA astrophysics theory program at the joint Center for Star Formation Studies at NASA-Ames Research Center, UC Berkeley, and UC Santa Cruz. REFERENCES Adams, F. C., Fatuzzo, M., 1996, ApJ, 464, 256 Adams, F. C., Myers, P. C., 2001, ApJ, in press (astro-ph/0102039) Ballesteros-Paredes, J., Vázquez-Semadeni, E., Scalo, J., 1999, ApJ, 515, 286 Bate, M. R., 2000, MNRAS, 314, 33 Bate, M. R., Bonnell, I. A, Price, N. M., 1995, MNRAS, 277, 362 Benz, W., 1990, in The Numerical Modeling of Nonlinear Stellar Pulsations, ed. J. R. Buchler (Dordrecht: Kluwer), 269 Bonnell, I. A., Bate, M. R., Clarke, C. J., & Pringle, J. E. 1997, MNRAS, 285, Clarke, C. J., Bonnell, I. A., Hillenbrand, L. A., 2000, in Protostars and Planets, eds. V. Mannings, A. P. Boss, & S. S. Russell, University of Arizona Press, Tucson, p. 151 Duquennoy, A., Mayor, M., 1991, A&A, 248, 485 Ebisuzaki, T., Makino, J., Fukushige, T., Taiji, M., Sugimoto, D., Ito, T., Okumura, S. K., 1993, PASJ, 45, 269 Elmegreen, B. G. 1993, ApJ, 419, L29 Elmegreen, B. G. 1993, MNRAS, 311, L5 Bergin, E. A., Goldsmith, P. F., Snell, R. L., Langer, W. D., 1997, ApJ, 428, 285 Heithausen, A., Bensch, F., Stutzki, J., Falgarone, E., Panis, J. F., 1998, A&A, 201 331, L65 Heitsch, F., Mac Low, M.-M., & Klessen, R. S. 2001, ApJ, 547, 280 Henriksen, R., André, P., Bontemps, S., 1997, A&A, 323, 549 Johnstone, D., Wilson, C. D., Moriarty-Schieven, G., Joncas, G., Smith, G., Gregersen, E., Fich, M., 2000, 545, 327 Klessen, R. S., 1997, MNRAS, 292, 11 Klessen, R. S., 2000, ApJ, 535, 869 Klessen, R. S., 2001, ApJ, 550, L77 (paper III) Klessen, R. S., Burkert, A., 2000, ApJS, 128, 287 (paper I) Klessen, R. S., Burkert, A., 2001, ApJ, 549, 386 (paper II) Klessen, R. S., Burkert, A., & Bate, M. R., 1998, ApJ, 501, L205 Klessen, R. S., Heitsch, F., Mac Low, M.-M., 2000, ApJ, 535, 887 Kramer, C., Stutzki, J., Rohrig, R., Corneliussen, U., 1998, A&A, 329, 249 Kroupa, P., 1995, MNRAS, 277, 1491 Kroupa, P., 2001, MNRAS, 322, 231 Kroupa, P., Tout, C. A., Gilmore, G., 1990, MNRAS, 244, 76 Lada, E., 1992, ApJ, 393, L25 Larson, R. B., 1973, MNRAS, 161, 133 Larson, R. B., 1998, MNRAS, 301, 569 Mac Low, M.-M., 1999, ApJ, 524, 169 Mac Low, M.-M., Klessen, R. S., Burkert, A., Smith, M. D., 1998, Phys. Rev. Lett., 80, 2754 Miller, G. E., Scalo, J. M., 1997, ApJS, 41, 513 Mizuno, A., Onishi, T., Yonekura, Y., Nagahama, T., Ogawa, H., Fukui, Y. 1995, ApJ, 445, L161 Motte, F., André, P., Neri, R., 1998, A&A, 336, 150 Myers, P. C., Evans, N. J. , Ohashi, N., 2000, in Protostars and Planets, eds. V. Mannings, A. P. Boss, & S. S. Russell, University of Arizona Press, Tucson, p. 217 Onishi, T. and Mizuno, A. and Kawamura, A. and Ogawa, H and Fukui, Y., 1996, ApJ, 465, 815 Ossenkopf, V., Mac Low, M.-M., 2001, A&A, submitted (astro-ph/0012247) Padoan, P., 1995, MNRAS, 277, 337 Padoan, P., Nordlund, A., 1999, ApJ, 526, 279 Padoan, P., Nordlund, A., 2001, ApJ, submitted (astro-ph/0011465) Padoan, P., Juvela, M., Goodman, A. A., Nordlund, A., 2001, ApJ, submitted (astro-ph/0011122) Salpeter, E. E., 1955, ApJ, 121, 161 Scalo, J., 1986, Fund. of Cos. Phys., 11, 1 Scalo, J., 1998, in The Stellar Initial Mass Function (38th Herstmonceux Conference), eds. G. Gilmore & D. Howell, ASP Conference Series, Vol. 142, p. 201 Scalo, J., Vázquez-Semadeni, E., Chappell, D., Passot, T., 1998, ApJ, 504, 835 Shu, F. H. 1977, ApJ, 214, 488 Shu, F. H., Adams, F. C., Lizano, S., 1987, ARA&A, 25, 23 Spaans, M., Silk, J., 2000, ApJ, 538, 115 Steinmetz, M., 1996, MNRAS, 278, 1005 Stutzki, J., Güsten, R., 1990, ApJ, 256, 513 Stone, J. M., Ostriker, E. C., Gammie, C. F., 1998, ApJ, 508, L99 10 Ralf S. Klessen Sugimoto, D., Chikada, Y., Makino, J., Ito, T., Ebisuzaki, T., Umemura, M., 1990, Nature, 345, 33 Williams, J. P., Blitz, L., McKee, C. F., 2000, in Protostars and Planets IV, eds. V. Mannings, A. P. Boss, & S. S. Russell, University of Arizona Press, Tucson, p. 97 Williams, J. P., De Geus, E. J., Blitz, L., 1994, ApJ, 428, 693 Wuchterl, G., Tscharnuter, W. M., 2000, A&A, submitted Testi, L., Sargent, A. I., 1998, ApJ, 506, L91 Testi, L., Palla, F., Natta, A., 1998, A&A, 342, 515 Zinnecker, H., 1984, MNRAS, 210, 43
astro-ph/0104118
1
0104
2001-04-06T13:30:20
Extremely red radio galaxies
[ "astro-ph" ]
At least half the radio galaxies at z>1 in the 7C Redshift Survey have extremely red colours (R-K>5), consistent with stellar populations which formed at high redshift (z>5). We discuss the implications of this for the evolution of massive galaxies in general and for the fraction of near-IR-selected EROs which host AGN, a result which is now being tested by deep, hard X-ray surveys. The conclusion is that many massive galaxies undergo at least two active phases: one at z~5 when the black hole and stellar bulge formed and another at z~1-2 when activity is triggered by an event such as an interaction or merger.
astro-ph
astro-ph
EXTREMELY RED RADIO GALAXIES Chris J. Willott, Steve Rawlings, Katherine M. Blundell Astrophysics, University of Oxford, UK Abstract At least half the radio galaxies at z > 1 in the 7C Redshift Survey have extremely red colours (R − K > 5), consistent with stellar popula- tions which formed at high redshift (z > ∼ 5). We discuss the implications of this for the evolution of massive galaxies in general and for the frac- tion of near-IR-selected EROs which host AGN, a result which is now being tested by deep, hard X-ray surveys. The conclusion is that many massive galaxies undergo at least two active phases: one at z ∼ 5 when the black hole and stellar bulge formed and another at z ∼ 1 − 2 when activity is triggered by an event such as an interaction or merger. Introduction Radio sources are known to reside in massive, luminous host galaxies. Therefore, they are an excellent way of selecting such galaxies out to high redshifts and tracing their evolution. The obscuring torus, which forms the basis of the unified schemes, blocks the non-stellar nuclear emission from our line-of-sight providing a much clearer view of the host galaxy properties than is the case for quasars. 3C radio galaxies at z > ∼ 0.6 have complex optical continuum and emission line structures aligned along their radio axes, which can generally be interpreted as due to recent (jet-induced) star-formation or non-thermal processes associated with the active nucleus (e.g. Best et al. 1998). Lacy et al. (1999a) have shown that the strength of the alignment effect decreases with decreasing radio luminosity such that fainter samples do not suffer from this problem. The 7C Redshift Survey (7CRS) is a low-radio frequency (151 MHz) flux-selected sample in three small patches of sky covering a total sky area of 0.022 sr. The flux-limit is 0.5 Jy -- a factor of 25× lower than the revised 3CR sample (Laing et al. 1983). We now have complete optical/near-infrared identifications for all the radio sources and > 90% spectroscopic redshifts. Further details of the survey are given in Willott 1 2 et al. (2001a) and Lacy et al. (1999b). Only 7 out of 76 sources in the 7C-I and 7C-II regions lack spectroscopic redshifts. For these sources we have obtained optical and near-IR photometry in order to constrain their redshifts using photometric redshift techniques. We find that most of these seven galaxies have very red colours. A full account of this work is given in Willott et al. (2001b). A flat cosmology with parameters ΩM = 0.3, ΩΛ = 0.7 and H0 = 70 km s−1Mpc−1 is assumed throughout. Results of photometric redshift fitting RIJ HK photometry was obtained for the seven sources without spec- troscopic redshifts. Six of these objects have extremely red colours (R − K > 5.5), similar to those expected from evolved stellar popu- lations at z > 1. These data were fit with a range of instantaneous burst model galaxies (Bruzual & Charlot in prep.) with redshift, age and reddening as free parameters. Best-fit models were found by search- ing for the minimum in the χ2 distribution. Fits with reduced χ2 < 1 were found for five objects and the remaining two had best-fit reduced χ2 < 2. An example of the SED-fitting is shown in Fig. 1. Example of model-fitting to the optical to near-IR SEDs of 7CRS ra- Figure 1. dio galaxies. The left panel shows the broad-band photometric data of 5C6.83 (7CB021111.3+303948; R > 24.5; K = 18.3) with best-fit model galaxies. The best fit unreddened model is shown as a solid line and the best-fit reddened model as a dotted line. Details of these fits are given in the bottom-right corner. The centre and right panels show how the χ2 of the fit depends upon redshift, age of the stellar population and reddening. The minimum in reduced χ2 (1.55) is shown as a cross. The best-fit redshifts for these radio galaxies ranges from z = 1.05 to z = 2.35 with typical uncertainties ∆z = ±0.3. Considering other factors such as the K − z relation and the lack of emission lines in their optical spectra, we believe that all the galaxies are likely to fall within the redshift range 1 < z < 2. The flatness of the SEDs from J to K strongly argues against z > 2.5. We know from near-IR spectroscopy that the colours are not strongly affected by emission line contamination, but in two cases there is a marginal emission line in the near-IR which Extremely red radio galaxies 3 could be Hα at z ≈ 1.5, consistent with the SED-fitting. Quite a wide range of parameters provide acceptable fits for many of the galaxies. The degeneracy of age and reddening appears to be the strongest cause of this as shown by the diagonal shape of contours in the age-reddening plane in Fig. 1 (from old and dust-free at top-left to young and dusty at bottom-right). In contrast there is little correlation between redshift and age in the contours and the best-fit redshifts are in most cases similar for both the reddened and unreddened cases. The colours of 7CRS radio galaxies Since all members of the 7CRS (regions I and II) have been securely identified, have spectroscopic or photometric redshifts and have R and K-band imaging (Willott et al. in prep), we can investigate the colour evolution of the radio galaxies. In Fig. 2 we plot the observed R − K colour against redshift for the 49 narrow-line radio galaxies and the 2 broad-line radio galaxies (the 23 quasars are not plotted because their magnitudes are clearly dominated by non-stellar emission). The solid curve is a model featuring an instantaneous starburst at redshift z = 5. At redshifts z ≤ 1 the colours of most of the radio galaxies are close to the model curve suggesting that these colours result from very old galaxy populations with little (unobscured) current star-formation. Deviations to bluer colours can be caused by a small amount of more recent star- formation or AGN-related processes like scattering of quasar light. Moving to z > ∼ 1 we find a marked increase in the scatter of the R − K colours of the radio galaxies. This is probably due to a combination of two effects. First, at redshifts higher than z ∼ 0.8, the observed R- band samples the rest-frame light below the 4000 A break. The ratio of fluxes below and above 4000 A is a very strong function of the amount of current or recent star-formation. Therefore small differences in the amount of recent star-formation will have a more dramatic effect on the observed colour at z > 0.8. Lilly & Longair (1984) showed that the optical/near-IR colours of 3CR galaxies at z > 1 are inconsistent with a no-evolution model. Their observed R − K ≈ 4 colours are similar to those of the bluest 7CRS radio galaxies. For the 3C objects these colours are best explained by recent star-formation in a few cases (Chambers & McCarthy 1990), but in most cases, as further evidenced by large optical polarizations (e.g. di Serego Alighieri et al. 1989), they are probably caused by an extra non-stellar rest-frame UV component, typically scattered light from the quasar nucleus (see Best et al. 1998). It is clear from Fig. 2 that, at z > 1, the bluer 7CRS galaxies are much more likely to have redshifts measured from optical spectroscopy. 4 Figure 2. Observed R − K colour as a function of redshift for radio galaxies in the 7C-I and 7C-II regions of the 7CRS. Triangles are narrow-line radio galaxies, crossed circles are broad-lined radio galaxies and filled squares are the seven 7CRS objects with photometric redshift estimates. The asterisks are the two red radio galaxies from the LBDS survey (Dunlop 1999). The solid line shows the evolution of the expected observed colour of a galaxy which formed in an instantaneous starburst at redshift z = 5 using the models of Bruzual & Charlot (in prep.). This model provides a good fit to the upper envelope of the galaxy colours up to at least redshift z = 2. This suggests that the bluer sources have stronger emission lines and are therefore good candidates for having non-stellar (AGN) and/or starburst components in the rest-frame UV. At z > 1, radio galaxies from the 6CE sample, which has a flux-limit in between those of 3CRR and the 7CRS, also tend to have strong emission lines and blue colours (Rawlings et al. 2001). Therefore the reason that such a high fraction of extremely red radio galaxies at high-redshift has not been seen in previous complete samples is due to their higher radio flux-limits selecting only higher luminosity z > 1 radio galaxies. Do the red colours indicate old galaxies? We have shown that at least half the z > 1 radio galaxies in the 7CRS have red colours (R −K > 5), consistent with those expected from evolved stellar populations at these redshifts. In addition, the fact that the bluer high-redshift radio galaxies have stronger emission lines, sug- gests that they too may have underlying host galaxies which are very red. Extremely red radio galaxies 5 But can we be certain that these colours are an indicator of age, rather than the effects of reddening by dust? The answer is that we cannot be completely certain in the absence of high signal-to-noise spectroscopy of stellar absorption features. Due to the faintness of these galaxies in the optical (I > ∼ 24), this will have to be performed in the near-infrared uti- lizing the low-background available with OH-suppression spectrographs on large telescopes. Instead, we have to look at circumstantial evidence, such as that attained for similar objects. Dunlop et al. (1999 and ref- erences therein) have studied two red radio galaxies at z ∼ 1.5 drawn from a faint radio sample. Deep Keck spectroscopy shows that the red colours of these galaxies are due to an old stellar population and not due to reddening by dust. The ages inferred are still controversial, although ages of ∼ 3 Gyr remain the best estimate (Nolan et al. 2001). For these ages observed at z = 1.5, the star-formation must have ceased at z ∼ 5. Implications for AGN activity in elliptical galaxies Although powerful radio galaxies are very rare objects, we can use our findings to predict the relationship between other AGN (both radio- loud and radio-quiet) and the near-infrared selected ERO population. Extrapolating down the radio luminosity function (Willott et al. 2001a) to less powerful radio galaxies, we find that if a similar fraction of these lower power radio galaxies have similar colours, they imply a space den- sity of red radio galaxies which is about 3% of the space density of near-IR selected EROs (Daddi et al. 2000). Hence we find that a small, but significant, fraction (∼ 3 %) of field EROs are likely to be hosting radio-loud AGN. However, it is well-known that luminous radio sources have limited lifetimes (∼ 108 years) which are much smaller than the Hubble time. The time elapsed between z = 2 and z = 1 is approximately 3.5 Gyr. Therefore if individual radio sources have lifetimes of only ∼ 108 years, then the number of galaxies undergoing radio activity during this period would be a factor of 30 greater than that observed. Hence all of the near-IR selected EROs could plausibly undergo such a period of radio activity. A caveat to this is that the typical lifetimes of weak radio sources such as those which would dominate the ERO population are not well-constrained and could have longer lifetimes of ∼ 1 Gyr. In such a case, only ∼ 10% of high-z EROs would undergo a period of radio activity at some point (see Willott et al. 2001b for more details). The hardness of the X-ray background requires that the space density of optically-obscured quasars exceeds that of optically-luminous quasars (e.g. Comastri et al. 1995), which in turn are well known to outnumber 6 radio-loud quasars by at least an order of magnitude. Many of the hard X-ray sources discovered in Chandra surveys have very red galaxy coun- terparts with weak or absent emission lines (Crawford et al. 2001; Cowie et al. 2001). These objects are likely to be the radio-quiet analogues of the 7CRS EROs discussed here. The hard X-ray properties of the ERO population will be investigated with XMM-Newton and Chandra sur- veys of ERO fields. In the HDF-North Caltech area, 4 out of 33 EROs (R − K > 5) have hard X-ray detections (Hornschemeier et al 2001). Assuming an elliptical galaxy fraction in the field ERO population of 70% (Moriondo et al. 2001; Stiavelli & Treu 2000) this corresponds to ∼ 20% of these ellipticals being observed to undergo a phase of AGN activity. Better statistics will accurately determine the duty cycle of AGN-activity in massive galaxies and provide constraints on quasar life- times (which are currently not well-constrained for radio-quiet quasars). It seems likely that rather than being an oddity, AGN activity is common in massive galaxies, not only during the high-redshift formation epoch, but also at a later stage once the major episode of star (and black hole) formation has long since ceased. References Best P.N., Longair M.S., Rottgering H.J.A., 1998, MNRAS, 295, 549 Chambers K.C., McCarthy P.J., 1990, ApJ, 354L, 9 Comastri A., Setti G., Zamorani G., Hasinger G., 1995, A&A, 296, 1 Cowie L.L., et al., 2001, ApJL, in press, astro-ph/0102306 Crawford C.S., Fabian A.C., Gandhi P., Wilman R.J., Johnstone R.M., 2001, MN- RAS, submitted, astro-ph/0005242 Daddi E., et al., 2000, A&A, 361, 535 di Serego Alighieri S., Fosbury R.A.E., Tadhunter C.N., Quinn P.J., 1989, Nature, 341, 307 Dunlop J.S., 1999, in The Most Distant Radio Galaxies, ed. P.N. Best, H.J.A. Rottgering, M.D. Lehnert, (KNAW Colloq.; Dordrecht: Kluwer), 14 Hornschemeier A.E., et al., 2001, ApJ, in press, astro-ph/0101494 Lacy M., Ridgway S.E., Wold M., Lilje P.B., Rawlings S., 1999a, MNRAS, 307, 420 Lacy M., et al., 1999b, MNRAS, 308, 1096 Laing R.A., Riley J.M., Longair M.S., 1983, MNRAS, 204, 151 Lilly S.J., Longair M.S., 1984, MNRAS, 211, 833 Moriondo G., Cimatti A., Daddi E., 2001, A&A, in press, astro-ph/0010335 Nolan L.A., Dunlop J.S., Jimenez R., Heavens A.F., 2001, MNRAS, submitted, astro- ph/0103450 Rawlings S., Eales S.A., Lacy M., 2001, MNRAS, 322, 523 Stiavelli M., Treu T., 2000, To appear in the proceedings of the conference "Galaxy Disks and Disk Galaxies", ASP Conf. series, eds. Funes and Corsini, astro-ph/0010100 Willott C.J., Rawlings S., Blundell K.M., Lacy M., Eales S.A., 2001a, MNRAS, 322, 536 Willott C.J., Rawlings S., Blundell K.M., 2001b, MNRAS, in press, astro-ph/0011082
astro-ph/9911182
1
9911
1999-11-10T20:56:28
Inversion of stellar statistics equation for the Galactic Bulge
[ "astro-ph" ]
A method based on Lucy (1974, AJ 79, 745) iterative algorithm is developed to invert the equation of stellar statistics for the Galactic bulge and is then applied to the K-band star counts from the Two-Micron Galactic Survey in a number of off-plane regions (10 deg.>|b|>2 deg., |l|<15 deg.). The top end of the K-band luminosity function is derived and the morphology of the stellar density function is fitted to triaxial ellipsoids, assuming a non-variable luminosity function within the bulge. The results, which have already been outlined by Lopez-Corredoira et al.(1997, MNRAS 292, L15), are shown in this paper with a full explanation of the steps of the inversion: the luminosity function shows a sharp decrease brighter than M_K=-8.0 mag when compared with the disc population; the bulge fits triaxial ellipsoids with the major axis in the Galactic plane at an angle with the line of sight to the Galactic centre of 12 deg. in the first quadrant; the axial ratios are 1:0.54:0.33, and the distance of the Sun from the centre of the triaxial ellipsoid is 7860 pc. The major-minor axial ratio of the ellipsoids is found not to be constant. However, the interpretation of this is controversial. An eccentricity of the true density-ellipsoid gradient and a population gradient are two possible explanations. The best fit for the stellar density, for 1300 pc<t<3000 pc, are calculated for both cases, assuming an ellipsoidal distribution with constant axial ratios, and when K_z is allowed to vary. From these, the total number of bulge stars is ~ 3 10^{10} or ~ 4 10^{10}, respectively.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000–000 (0000) Printed 2 March 2018 (MN LATEX style file v1.4) Inversion of stellar statistics equation for the Galactic Bulge M. L´opez-Corredoira ⋆1, P. L. Hammersley1, F. Garz´on1,2, E. Simonneau3, T. J. Mahoney1 1 Instituto de Astrof´ısica de Canarias, E-38200 La Laguna, Tenerife, Spain 2 Departamento de Astrof´ısica. Universidad de La Laguna, E-38204 La Laguna, Tenerife, Spain 3 Institut d'Astrophysique de Paris, F-75014 Paris, France Accepted xxxx. Received xxxx; in original form xxxx ABSTRACT A method based on Lucy (1974) iterative algorithm is developed to invert the equation of stellar statistics for the Galactic bulge and is then applied to the K-band star counts from the Two-Micron Galactic Survey in a number of off-plane regions (10◦ > b > 2◦, l < 15◦). The top end of the K-band luminosity function is derived and the morphology of the stellar density function is fitted to triaxial ellipsoids, assuming a non-variable luminosity function within the bulge. The results, which have already been outlined by L´opez-Corredoira et al. (1997b), are shown in this paper with a full explanation of the steps of the inversion: the luminosity function shows a sharp decrease brighter than MK = −8.0 mag when compared with the disc population; the bulge fits triaxial ellipsoids with the major axis in the Galactic plane at an angle with the line of sight to the Galactic centre of 12◦ in the first quadrant; the axial ratios are 1:0.54:0.33, and the distance of the Sun from the centre of the triaxial ellipsoid is 7860 pc. The major–minor axial ratio of the ellipsoids is found not to be constant, the best fit to the gradient being Kz = (8.4 ± 1.7) × exp (−t/(2000 ± 920) pc), where t is the distance along the major axis of the ellipsoid in parsecs. However, the interpretation of this is controversial. An eccentricity of the true density-ellipsoid gradient and a population gradient are two possible explanations. The best fit for the stellar density, for 1300 pc < t < 3000 pc, are calculated for both cases, assuming an ellipsoidal distribution with constant axial ratios, and when Kz is allowed to vary. From these, the total number of bulge stars is ∼ 3 × 1010 or ∼ 4 × 1010, respectively. Key words: Galaxy: structure - infrared: stars - Galaxy: stellar content - stellar statistics 1 INTRODUCTION This paper examines two aspects of the bulge: the luminosity function for the brightest stars in the K (2.2 µm) band and the density distribution of these stars. There are many aspects of the bulge of the Galaxy that are still unknown, mainly because of the high extinction due to interstellar gas and dust. One of these unknowns is the near-infrared luminosity function, which has been princi- pally derived from observations in Baade's Window (Frogel & Whitford 1987; Davidge 1991; De Poy et al. 1993; Ruelas- Mayorga & Noriega-Mendoza 1995; Tiede et al. 1995). Gould (1997) and Holtzman et al. (1998) have used the Hubble ⋆ Electronic mail: [email protected]. c(cid:13) 0000 RAS Space Telescope to study the V and I luminosity functions. However, extrapolations from Baade's and other clear win- dows to the whole bulge may not be appropriate, in partic- ular because these are 'special' regions. Furthermore, these regions are very small, containing relatively few stars, so they give very poor statistics at the brighter magnitudes. This bright end of the luminosity function is very important in order to determine the age of the population, for instance. Many authors have found non-axisymmetry in the Galactic bulge† (Feast & Whitelock 1990) through the anal- ysis of star counts (Nakada et al. 1991; Weinberg 1992; † Some authors call it the "bar" instead of the bulge. See, for instance, Gerhard, Binney & Zhao (1998). 2 L´opez-Corredoira et al. Whitelock et al. 1991; Stanek et al. 1994, 1996; Wo´zniak & Stanek 1996) or integrated flux maps (Blitz & Spergel 1991; Weiland et al. 1994; Dwek et al. 1995; Sevenster 1996). This asymmetry has a negligible out-of-plane tilt (Weiland et al. 1994) and gives more counts in positive than in neg- ative galactic longitudes. However, other authors (Ibata & Gilmore 1995; Minniti 1996) claim that axisymmetry is suit- able. Besides the discussion about whether there is triaxial- ity or not, the actual shape and inclination of the bulge is also under debate with currently no clear agreement among different authors. Traditionally, star counts have been interpreted by fit- ting parameters to the functions. An assumption of an a priori shape of the bulge is made, along with the character- istics of its population. Free parameters are then fitted to the data and the model is obtained. This is the usual way of ex- tracting information concerning the different components of the Galaxy from star counts (Bahcall & Soneira 1980; Buser & Kaeser 1983; Prichet 1983; Gilmore 1984; Robin & Cr´ez´e 1986; Ruelas-Mayorga 1991; Wainscoat et al. 1992; Ortiz & L´epine 1993). The number of possible parameters to fit is limited to a priori assumptions about the shape (ellipsoidal, etc.) needed. In general, surface brightness maps are also interpreted by fitting parameters (Dwek et al. 1995; Freudenreich 1998). However, although these maps cover large areas, a brief ex- amination of the equations shows that they give no infor- mation on the luminosity functions. Therefore, when mak- ing the fit to the bulge, the number of free parameters is very small and applies only to the density function. Even in Freudenreich (1998), where in total some 30 parameters are used, only a very few of these apply to the bulge and only a very few parameters are solved at one go. In this paper we examine the TMGS star counts be- tween mK=4 and 9 mag in 71 regions across the bulge. The counts for these regions are shown in Hammersley et al. (1999), where a qualitative discussion on the counts is presented. It is shown that the counts are highly asymmet- ric in longitude when compared with the predictions of a symmetric model. Clearly, there is a relation between surface brightness and star counts as one is the integral of the other, but they are not the same or even similar and cannot be handled in the same manner. One way of looking at the difference be- tween the two is to consider that at a single position a surface brightness map gives just a single value whilst star counts give a counts. vs magnitude plot. From this plot alone it is possible to determine things about the structure. Therefore, while star counts and surface brightness maps are clearly related, they behave very differently. Many authors, includ- ing ourselves, have already used the fitting approach to look at the surface brightness COBE-DIRBE data (we note that Binney et al. 1997 have tried inversions on the surface bright- ness maps), but this is not the best for star counts in the present situation. One of the major advantages in analysing star counts as opposed to surface brightness maps is that the magnitude range can be limited in order to highlight the features of interest. This is of particular value when looking for triaxi- ality because if one region is significantly closer than another then, simply from the inverse square dependence with the distance, the sources from the further region are not de- tected until a fainter magnitude. Hammersley et al. (1999) show that in the TMGS star counts the size of the asym- metry amounts to some 50% of the bulge counts in some magnitude ranges, in the COBE-DIRBE maps the asymme- try is far less. For this reason analysis of 2-µm star counts in a certain magnitude range will be far more sensitive in de- termining the triaxiality of the bulge than surface brightness maps. Whilst large-area star counts, as used here, contain far more information than surface brightness maps they are in- trinsically far more difficult to analyse. A priori, neither is known and furthermore there is no reason to believe that the luminosity function (LF) is a simple analytical expression. Therefore, whereas fitting a surface brightness map there will only be a few free parameters, this is not the case for star counts. In this case the number of free parameters would rise unmanageably and so we would be forced to adopt a pri- ori assumptions on the LF and density functions with the severe risk that the final result is dependent on these initial assumptions. We have therefore chosen a different approach, that of direct inversion. Assumptions on the shape of the solution- functions (in this case, these are the luminosity function and the density of stars) are not made but instead come directly from the data by means of an "inversion" technique. Once the solutions for the functions are produced by the inversion, they are compared a posteriori to some known analytical expression (for instance, an ellipsoidal shape for the bulge isodensity contours) and, afterwards, fitted to them. This method allows all possible solutions to be examined, rather than solely that of the initial assumption and the only fitting are density contours to a density map. No attempt need to be made to fit a density function to the star counts. Since the first decades of this century, attempts have been made to invert the star-count equation (eq. 1). How- ever, problems such as excessive patchiness of extinction in optical star counts or instabilities of an ill-posed problem in the mathematical technique of inversion, hindered the development of the technique. In this paper the extinction problems are ameliorated by using the near-infrared K band and the instabilities by using a statistical iterative algorithm of inversion (Lucy 1974). A full explanation of the inversion is developed in this paper (with the core in §4) whose re- sults have already been outlined by L´opez-Corredoira et al. (1997b). 2 NEAR-INFRARED DATA K-band star counts were taken from the Two Micron Galac- tic Survey (TMGS; Garz´on et al. 1993, 1996), which covers about 350 deg2 of sky and has detected some 700000 stars in or near the Galactic plane. This survey provides K-band observations of several regions that cross the Galactic plane, in the areas −5◦ < l < 35◦, b ≤ 15◦ and 35◦ < l < 180◦, b ≤ 5◦. Regions from three strips of constant declination are used (Table 1) In this study of the bulge. More specif- ically, 71 regions were selected from those strips in off- plane regions, but not too far from the Galactic centre (10◦ > b > 2◦, l < 15◦). Each region has an area on the sky between 0.4 and 1.9 deg2. The chosen regions are listed c(cid:13) 0000 RAS, MNRAS 000, 000–000 Inversion ... Bulge 3 Table 1. Constant-declination TMGS strip used in this paper. δcentral(J2000) Cut in the Galactic plane Strip width (∆δ) −29◦43′32′′ −22◦26′40′′ −15◦33′24′′ (deg) l = −0.9 l = 7.5 l = 15.4 (deg) 2.51 1.63 0.78 can be seen in the figure, the counts are nearly the same with or without correction. That confusion is not significant for the areas chosen can also be seen in the figures in Hammer- sley et al. (1999) where the TMGS star counts are directly compared with the W92 model counts (Cohen 1994). Taking into account that the correction is based on an extrapola- tion and the changes are minor when compared to the other sources of error, it is preferable to avoid any correction and use the original counts. 3 THE STELLAR STATISTICS EQUATION FOR THE BULGE 3.1 Cumulative star counts For each of the 71 regions centred on galactic coordinates (l, b)i, where i is the field number, the cumulative star counts observed in a filter K, NK , up to a magnitude mK in a given region of solid angle ω is the sum of the stars over the beam with such an apparent magnitude (Bahcall 1986). Assuming a luminosity function which does not vary with the spatial position for each Galactic component c, this is ΦK,c(mK + 5 − 5 log10 r − aK(r)) NK (mK)ωXc Z ∞ 0 × Dc(r)r2dr, where ΦK,c(MK ) =Z MK −∞ φK,c(M )dM., (1) (2) φK,c is the normalized luminosity function for the K band in the component c; Dc is the density function in the com- ponent c and aK(r) is the extinction along the line of sight for the K band. 3.2 Extinction If the star counts, NK , for eq. (1), and the luminosity func- tion, φK,c, are known then the densities and the extinction would be the unknown functions. The extinction can be sep- arated from the last integral equation by means of a suitable change of variable (Bok 1937; Trumpler & Weaver 1953; Mi- halas & Binney 1981, ch. 4): ρK = 100.2aK (r)r, ∆c,K[ρK (r)] = Dc(r) which transforms the stellar statistics equation into (cid:0)1 + 0.2(ln 10)r daK (r) dr (cid:1) 100.6aK (r) (3) (4) , Figure 1. Comparison of cumulated star counts without con- fusion correction and those corrected according to the method explained in L´opez-Corredoira et al. (1997a) with a linear ex- trapolation of the differential star counts over magnitude 9.4. in Table 2. There is an overlap in the neighbouring regions such that some stars fall into two regions. The total covered area of sky covered is 75 deg2. This area is far greater than that used in Baade's window or any of the other low extinc- tion region and hence provides much better statistics for the top end of the bulge LF. The chosen regions contain principally bulge and disc stars. The area near the Galactic plane was not used in order to avoid components which belong neither to the bulge nor to the disc (e.g. spiral arms) and the high and variable extinction. The outer limits were set so that the bulge-to- disc stellar ratio was still acceptable, i.e. so that there were sufficient bulge stars in comparison with disc stars to make the study of the bugle meaningful.. The survey is complete between the magnitude limits mK = 4.0 mag and mK ≈ 9.2 mag, except for the regions very near the Galactic centre where source confusion re- duced the faint limit by about half a magnitude, although the detection limiting magnitude of the survey is in excess of 10 mag. Hence, inversion will be applied up to mK = 8.6 mag for the regions of the strip with declination −30◦ and up to mK = 9.0 mag for the remaining cases. Figure 2 shows cumulative star counts, N , for the three strips up to mK = 9 mag as a function of b (l also varies, as can be seen in Fig. 9). Within this range of magnitudes, confusion effects are negligible. This was determined from the application of the method explained by L´opez-Corredoira et al. (1997a) by as- suming an extrapolation to fainter magnitudes (Fig. 1). As c(cid:13) 0000 RAS, MNRAS 000, 000–000 4 L´opez-Corredoira et al. Table 2. The regions whose star counts are used to invert and extract information about the bulge. l b (deg) (deg) Area (deg2) l b (deg) (deg) Area (deg2) l b (deg) (deg) Area (deg2) –6.3 –5.7 –5.2 –4.7 –4.2 –3.7 –3.2 –2.7 –2.6 –2.5 –2.4 –2.3 –2.3 0.3 0.4 0.5 0.6 0.7 0.7 0.8 0.9 1.3 1.8 2.2 7.8 7.1 6.4 5.7 5.0 4.3 3.5 2.8 2.7 2.5 2.4 2.2 2.1 –2.0 –2.2 –2.3 –2.5 –2.6 –2.8 –2.9 –3.1 –3.9 –4.6 –5.4 0.4 0.8 1.3 1.8 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 1.9 2.6 3.0 3.4 3.8 4.2 1.3 1.8 2.3 2.9 3.4 3.9 4.4 4.9 5.4 5.8 5.9 6.0 6.1 6.2 6.3 8.7 8.8 8.9 8.9 –6.1 –6.9 –7.7 –8.5 –9.2 9.9 9.1 8.4 7.6 6.8 6.0 5.3 4.5 3.7 2.9 2.7 2.6 2.4 2.3 2.1 –2.1 –2.2 –2.4 –2.6 1.9 1.9 1.8 1.8 1.8 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 9.0 9.1 9.2 9.6 10.1 10.5 10.9 11.3 11.7 12.2 12.6 9.7 10.2 10.7 11.2 11.7 12.2 12.6 13.1 13.6 14.1 14.2 14.3 –2.7 –2.9 –3.0 –3.9 –4.7 –5.5 –6.3 –7.1 –8.0 –8.8 –9.6 9.9 9.1 8.3 7.4 6.6 5.8 4.9 4.1 3.3 2.4 2.2 2.1 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 1.4 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 NK (mK) = ωXc Z ∞ 0 × ∆c,K(ρK)ρ2 KdρK . ΦK,c(mK + 5 − 5 log10 ρK) (5) The functions ∆c,K(ρK) do not have a direct physical meaning but are fictitious densities as a function of a fic- titious distance which coincides with the real distance only when there is no extinction (see Calbet et al. 1995). For the extinction we have followed Wainscoat et al. (1992, hereafter W92), who assume that the extinction has an exponential distribution with the same scale length as the old disc, 3.5 kpc, and a scale height of 100 pc. This is normalized to give AK = daK/dr = 0.07 mag kpc−1 in the solar neighbourhood. Although this model is crude it is suf- ficient for our purposes. As the areas of interest are off the plane, the extinction in the direction of the bulge sources is between 0.05 to 0.5 mag at K (ten times lower than in V ). The evidence from the 2.2-µm surface brightness maps is that there are off-plane clouds, but these are isolated so if a strip did cross a cloud it would affect only one or two regions which would have a minor effect on the final result. In fact, Hammersley et al. (1999) show that in the regions chosen there are no major dips in the counts and hence no isolated clouds. Furthermore, in this paper there is a dis- cussion on the IR extinction in the plane and comparison is made with the W92 model, which uses the above model for the extinction. It is shown that in the solar neighbourhood this model works well and remains valid to a galactocentric distance of about 4 kpc where the molecular ring is situated. Inside the ring the extinction is then over estimated. How- ever, it should be noted that for the lines of sight used here the majority of the extinction occurs in the first few kpc, i.e. while the line of sight is close to the Galactic plane. There- fore, the extra extinction added by the model in the inner galaxy is a small proportion of the total extinction along the line of sight, which is in turn already small. This effect is clearly demonstrated by Hammersley et al (1999) for the l = 7◦ strip where the effect of the overestimated extinction can only be seen within 0.5◦ of the plane. Another possible cause for concern could be if there was a general asymmetry in the extinction, either from above to below the plane or between positive and negative longitudes. However, it must be noted that the analysis of Freuden- rich (1998) of the COBE-DIRBE surface brightness maps shows no such asymmetry. Furthermore, Hammersley et al. (1999) have analysed the asymmetry in the TMGS bulge star counts and show that the form is not consistent with the asymmetry being caused by extinction. Therefore, although the extinction model is crude it is valid for the purpose used here. So, from aK , the relationship is obtained between ∆ -the fictitious density- and D -the real density- for each component, using eq. (3); therefore eq. (5) will be used hereafter. 3.3 Subtraction of the disc The components cannot all be solved simultaneously and the inversion of eq. (5) can only be solved when the number of components, c, is restricted to one. It will be assumed that, in the chosen regions, the con- tribution to the star counts will be primarily from the disc and bulge. In order to isolate the bulge component, there- fore, the contribution of the disc must be subtracted from the total counts for each region. The model of the disc coded by us was based on W92, c(cid:13) 0000 RAS, MNRAS 000, 000–000 Inversion ... Bulge 5 Figure 2. N (mK = 9.0 mag) along the three strips that are used with constant declinations: δ = −30◦, which cuts the plane at l = −1◦; δ = −22◦, which cuts the plane at l = 7◦; and δ = −16◦, which cuts the plane at l = 15◦. which follows Bahcall & Soneira (1980). It has been used because it provides a good fit to the TMGS counts in the region where the disc dominates (Cohen 1994b; Hammersley et al. 1999). The W92 model was revised by Cohen (1994a) but this does not significantly alter the form of the disc in the areas of interest. Three examples of those fits are shown in Fig. 3, in regions where the disc is isolated (note that the regions used in these plots are different from the regions used for the inversion specified in §2). A more detailed com- parison of the W92 model and the TMGS is presented by Hammersley et al. (1999), who examine some 300 square degrees of sky. Hence, by extrapolation, it is expected that this disc model will adequately reflect the disc components along the lines of sight used in this paper. Initially, it was also expected that the W92 model would give an adequate fit for the bulge counts; this, however, was not the case as can be clearly seen in Hammersley et al. (1999). 3.4 Fredholm integral equations of the first kind Once the disc star counts are subtracted, a Fredholm integral equation of the first kind is derived (see Trumpler & Weaver 1953): NK,bulge(mK) = NK (mK) − NK,disc(mK) c(cid:13) 0000 RAS, MNRAS 000, 000–000 0 KdρK, ΦK,bulge(mK + 5 − 5 log10 ρK)∆bulge,K (ρK) = ωZ ∞ × ρ2 where ∆ is the unknown function and Φ is the kernel of the integral equation. (6) When Φ is the unknown function instead of ∆, then a new change of variable can be made: MK = mK + 5 − 5 log10 ρK, and a new Fredholm equation of the first kind is obtained: NK (mK) = 200(ln 10)10 3mK 5 ∆K (10 5+mK −MK 5 )10 −3MK 5 ΦK (MK )dMK . (7) ×Z ∞ −∞ In this case, the kernel is ∆ instead of Φ. Any method of inverting eq. (6) is also applicable to this integral equation (7). 4 INVERSION OF THE STELLAR STATISTICS EQUATION The inversion of integral equations such as (6) or (7) is ill- conditioned. Typical analytical methods for solving these equations (see Bal´azs 1995) cannot achieve a good solution 6 L´opez-Corredoira et al. a) c) Figure 3. Differential star counts, the derivative of the cumu- lative star counts. Rhombi are TMGS data. Lines represent the W92 model: the solid line stands for counts for all components; the dotted line stands for disc counts; long-dashed line for spiral arms; short-dashed and dotted line for the ring; shot-dashed line for the bulge; long-dashed and dotted line for the halo. In these cases - a), b) and c) - disc and total counts are nearly coincident because the disc gives the most part of the stars. In eq. (6), ∆ is the unknown function, and the kernel is Φ, which depends on the apparent magnitude conditioned to the fictitious distance ρ. The fictitious density ∆ can also be understood in terms of a probability density (the probability of finding a star with fictitious distance ρ). Thus, eq. (6) can be rewritten as (hereafter, the notation for component or passband will be dropped) N (m) =Z ∞ 0 ∆(ρ)P (mρ)dρ, (8) where P (mρ) = ρ2Φ(m + 5 − 5 log10 ρ). The inverse conditioned probability, i.e. the probability of star being at a fictitious distance ρ, once its apparent mag- nitude m is known, is given by Bayes' theorem: (9) ∆(ρ)P (mρ) ∆(x)P (mx)dx . Q(ρm) = 0 R ∞ From the definition of conditioned probability, (10) (11) (12) (14) b) because of the sensitivity of the kernel to the the noise of the counts (see, for instance, Craig & Brown 1986, ch. 5). Since the functions in these equations have a stochastic rather than analytical interpretation, it is to be expected that statistical inversion algorithms will be more robust. This is confirmed by several authors, for instance Turchin et al. (1971), Jupp et al. (1975), Bal´azs (1995). From among these statistical methods, we have selected Lucy's algorithm (Lucy 1974; Turchin et al. 1971; Bal´azs 1995), an iterative method, the key to which is the inter- pretation of the kernel as a conditioned probability and the application of Bayes' theorem‡. ∆(ρ)P (mρ) = N (m)Q(ρm), and, hence, we get directly: mmin ∆(ρ) = R mmax R mmax mmin dmN (m)Q(ρm) . dmP (mρ) N obs (m) mmin ∆r+1(ρ) = ∆r(ρ)R mmax R mmax ∆r(x)P (mx)dx. N r(m) =Z ∞ mmin 0 Equations (12) and (10) together lead to an iterative method§ of obtaining the unknown function ∆(ρ): N r (m) P (mρ)dm P (mρ)dm , (13) where N obs represents the observed cumulative counts and ‡ Bayesian methods have multiple applications in astrophysics. Inversion problems are particular cases of these applications (Loredo 1990). § For the numerical calculation of these integrals ρ is placed into discrete logarithmic intervals (the (m, log π) method; Mihalas & Binney 1981, ch. 4) in such a way that log10 ρK is regularly spaced. c(cid:13) 0000 RAS, MNRAS 000, 000–000 This development is more general than Lucy's. Lucy's algorithm (Lucy 1974)'s algorithm was expressed for cases P (mρ) = 1, which is not true in the case dis- cussed here because the range of magnitudes is limited. The need for the denominator in eq. (13) was already recognized by Scoville et al. (1983). with R mmax mmin The iteration converges when N r = N obs, i.e. when ∆r+1 = ∆r. The first iterations produce a result which is close to the final answer, with the subsequent iterations giv- ing only small corrections. This algorithm has a number of good properties (Lucy 1974, 1994): both the luminosity function and the density are defined as being positive, the likelihood increases with the number of iterations, the method is insensitive to high frequency noise in N obs, etc. 4.1 Stopping criteria for the iterative process and initial trial solution From Lucy (1994), the appropriate moment at which to stop this kind of iterative process is when the curvature of the trajectory in the H–S diagram is a minimum. H and entropy (S) are defined by: N obs j ln N r j H =Xj and ∆r i ln ∆r i ∆0 i , S = −Xi (15) (16) respectively, and the curvature in the H–S diagram is: κ = S ′H ′′ − H ′S ′′ (S ′2 + H ′2)3/2 , (17) where the derivatives are with respect to the number of it- erations, and the sums over i and j correspond to discrete values of the ρ and m integrals respectively. Tests were carried out on the data set using this crite- rion (see an example in Fig. 4). In general there is a mini- mum after three iterations, corresponding to a non-relaxed state of the process. Afterwards, κ is increase up to around 10 iterations, where it then falls off again to a minimum, and then increases again. Apart from first minimum at 3 iterations, the most relevant minimum seems to be that at around 10 iterations. However, this criterion is not very accurate for the nois- iest cases and on occasions the last iteration may not be the most appropriate one to end at. Occasionally, it stops too early and therefore hinders the extraction of further infor- mation that could be exploited. Therefore, the following criteria are adopted for ending the iterations: (i) The number of iterations must be greater than 10 and smaller than 10000. The process will always be stopped when the number of iterations exceeds 10000. The ∆r variations are too small after 10000 iterations, so no more are made. (ii) For fewer than 1000 iterations, the iterative process is stopped when the solution is within the noise, i.e. when the average over m of the distance between N r(m) and N obs(m) c(cid:13) 0000 RAS, MNRAS 000, 000–000 Inversion ... Bulge 7 0.0020 0.0015 κ 0.0010 0.0005 0.0000 0 10 20 30 iter. Figure 4. Curvature versus iteration number in an inversion case. is less than the average over m of a random noise with Gaus- sian distribution of N obs(m) with σm = S(N obs(m)), the Poissonian noise of N obs(m). This last point will be clarified and the numerical algo- rithm to be used explained in what follows. N r i (the subindex i stands for the discrete value of m) is at si σs from N obs , i.e. si = N r (18) . i i − N obs i S(N obs ) i The normalized probability of a point at distance si σs from its real value is pi(si) = erf(si), (19) du is the error function. Thus, since the pi distribution is nearly uniform between 0 and 1, then the si distribution follows 0 e−u2 where erf(x) = (2/√π)R x Pn ≈Z 1 i=1 pi(si)2 p2 i dpi = n 0 1 3 . (20) 'Nearly' because it is exact when n → ∞, and there are some fluctuations when n is not too large. Thus, within the noise means that i=1 pi(si)2 Pn n < 1 3 , (21) and this is second stopping criterion. The sum of p2 i is calculated instead of the sum of pi be- cause the difference distribution is not exactly Gaussian and a power of pi gives a higher weighting to the large deviations (larger than 1–2 σ). In any case, this is only an approximate criterion. The final solution does not depend on the initial trial solution, N 1, when the number iterations is high enough. However, N r may approach N obs in a different way depend- ing on the initial trial solution when the noise of the counts is high, because the process is stopped after a few iterations 8 L´opez-Corredoira et al. which will give slightly different solutions. In order to avoid this influence for the noisiest data to be inverted¶, the trial solution was fed back with the smoothed result of the pre- vious inversion and inverted again. As will be discussed in §4.4, three inversions are made. In the second and the third iterations the trial solutions are fed back with the previ- ous outcome, once it has been fitted to a smooth analytical function (in this case ellipsoids, as seen in §6). 4.2 Distance range As the case described here is the application of the method to the bulge of our Galaxy, the numerical calculation of the distance integral are carried out over 2000 pc< ρK < 30000 pc, as all of the stars are known to be contained within this distance. The real distance, r, is somewhat lower than ρ (see eq. 3), but the difference is small for low-extinction regions such as those used here. It noted that, following numerical experiments with Lucy's algorithm, the minimum distance has to be kept within tolerable limits. Spurious fluctuations arise when small distances are included. This is related to the propor- tionality of the kernel to ρ2, so that large variations in the density at small distances do not significantly change the number of counts. A similar problem arise for the maximum distance to which the sources can be distributed. If the maximum limit is too large then a spuriously high density might appear at large distances. The reason is that very distant stars should be very luminous to be observed and, since the luminosity function for very luminous stars is very small, any sources placed at a large distance will lead to a high density at that distance. The application of this method to the bulge does not lead to problems since the distance range is known to be limited: the Galaxy has a boundary and the number of bulge stars in the solar neighbourhood is negligible. Nevertheless, it should be noted that care should be taken before applying this method to other Galactic components. For instance, it is possible that inverting the counts to obtain the Galactic disc density could encounter the above problems. 4.3 Example of application Inversion of the stellar statistics equation has been discussed by many authors, much more often in theory than in prac- tice, and doubt has been cast on the viability of such an inversion. It has even even said that as the solution is non- unique (Gilmore 1989), which would lead to instability in the inversion. Except for some particular kernel functions (Craig & Brown 1986; ch. 4) this is not in fact the case, as we shall attempt to demonstrate here. The question of uniqueness is important only from a theoretical standpoint. In practice, the only relevant issue is whether the method is able to obtain a solution close to the real one when the counts are affected by noise, which always produces devi- ations from the real solution. The important thing is that ¶ Very noisy data are eliminated. In this case, the 37 least noisy regions out of 71 are used when the density is the unknown function. ) ρ ( ∆ 1.5 1.0 0.5 0.0 0 final initial 5000 10000 ρ 15000 20000 a) Figure 5. Recovery of the theoretical through the inversion process. Three cases: a), b), c). luminosity function ) ρ ( ∆ 0.50 0.40 0.30 0.20 0.10 0.00 0 final initial 5000 10000 ρ 15000 20000 b) this solution be not very far from the true solution. That the solution is not unique need not be important when all solutions be close to each other. In order to test the reliability of the method, a num- ber of simulations were made. A luminosity function and a fictitious density function were constructed. The cumulative count per square degree, N (m), were then calculated by in- tegrating eq. (6). A random noise with a Gaussian distribu- tion is added to each bin. The cumulative counts with noise are then represented by N obs(m). When Lucy's algorithm is applied, with the same luminosity function and ∆1(ρ) = 1 (the choice of the trial initial solution does not affect the outcome), the results shown in Fig. 5 are obtained. The inversion is not perfect since it is affected by the noise, but the results are fairly good. There is a "hump" in the first case at short distances, and a large increase in the density at large distances in the second case. However, it should be noted that the hump is at the 10% level of c(cid:13) 0000 RAS, MNRAS 000, 000–000 ) ρ ( ∆ 1.5 1.0 0.5 0.0 0 final initial 10000 20000 ρ 30000 40000 c) the primary peak, which is located very close to the correct distance of 10000 pc. Similarly, the excess at large distances is at the level of only a few percent of the peak sources. Sensitivity to noise is higher for distances less than 7000 pc or greater than 15000 pc, as explained in §4.2, and this is reproduced in the experiments. When the method is asked to recover two peaks, the inversion gives poorer results. However, were the number of iterations increased beyond the 10000 limit, then the second peak in Fig. 5 c) would rise and become closer to the orig- inal. Again, however, both peaks are correctly located and the total number of sources in each peaks is very close to the original. Apart from these details, the general shape of the peaks is recovered. Other numerical experiments were performed with similar results. The bulge is a single-peaked structure so the proposed stopping criteria are sufficient. Since noise is random, the composition of the three-dimensional densities from the in- version for different regions (l, b) will attenuate the average deviations. Application to equation (7), instead of (6), deserves sim- ilar considerations. 4.4 Method of deriving both the luminosity function and the density The equations (6) and (7) can be solved for either the lu- minosity function or the density function, but not for both simultaneously for each region. Since both functions, ∆ and Φ, are of interest but accurate information is not available for either of them, the following method was used. To begin with, a first order approximation for the den- sity was assumed. It was taken from the axisymmetric W92 model. A simple comparison showed that the W92 luminos- ity functions suggested that there were possible problems with the brightest sources, although the density function did give a reasonable starting point. Therefore, it was de- cided to solve first for the average luminosity function using the W92 bulge density. With this density distribution, eq. (7) is inverted by means of Lucy's algorithm to provide the luminosity func- c(cid:13) 0000 RAS, MNRAS 000, 000–000 Inversion ... Bulge 9 tion for each of the regions (l, b) in Table 2. The weighted average of all luminosity functions was then calculated. We have made the assumption that the bulge luminosity function is independent of position. This assumption is sus- pect (see Frogel 1988, Section 3) since the observed metal- licity gradient might affect the luminosity of the AGB stars, although not the non-variable M-giants whose bolometric luminosity function is nearly independent of the latitude (Frogel et al. 1990). Some authors claim that there is a pop- ulation gradient (Frogel 1990; Houdashelt 1996; Frogel et al. 1999), while others do not (Tyson & Rich, 1993, show that there is no metallicity gradient up to 10◦ out of the plane; Ibata & Gilmore, 1995, argue that there is no detectable abundance gradient in the Galactic bulge over the galacto- centric range from 500 to 3500 pc). While the assumption may not be strictly true, it is nevertheless a useful approxi- mation in deriving mean properties of the bulge. With this averaged luminosity function, eq. (5) was in- verted to derive a new density distribution by means of Lucy's algorithm for each region. In this step the 37 regions with the highest counts were used, as the determination of the density is more sensitive to noise. The inversion of the luminosity function is more stable because the density distribution is sharply peaked and so the kernel in eq. (7) behaves almost as a Dirac delta func- tion. Hence, the shape of the density distribution does not significantly affect the shape of the luminosity function. The new density was then used to improve the lumi- nosity function, etc. The whole process was iterated three times, which was enough for the results to stabilize as can be seen in Fig. 6: it is seen how the result of the third iteration is very close to the first, i.e. stabilization is reached in the first iterations. This small variation in successive iterations is really a convergence to the solution since, as is shown in §6.3.4 and §6.4.1, the counts are approximately recovered when we project the bulge obtained from the inversion. The functions of interest are φ, the derivative of Φ, and D, related to ∆ by the change of variable expressed in eqs. (3) and (4). 5 THE TOP END OF THE K LUMINOSITY FUNCTION After three iterations the luminosity function was nearly independent of the position (l, b)i, stable and hardly changed from the solution of the second iteration. Compare the first three iterations in Figure 6. In fact, even the first iteration came close to the final solution. The obtained luminosity function is shown in Fig. 7 and in table 3. The derivative, φ, of Φ(MK ), from eqs. (7) and (2) is the normalized probability of having absolute magnitude MK per unit absolute magnitude. Figure 7 shows that for −10 mag < MK < −8 mag the bulge luminosity function is significantly lower than that of the disc (Eaton et al. 1984). Hence, the density of very bright stars in the bulge is much less than in the disc. Fainter than MK = −8 mag the luminosity functions of the disc and the bulge coincide, in agreement with Gould (1997). The lumi- nosity function for -10 mag < MK < −8 mag is significantly below the synthesized luminosity function assumed by W92 for the bulge in their model of the Galaxy (this can also be 10 L´opez-Corredoira et al. iter. 1 iter. 2 iter. 3 −4.0 −5.0 −6.0 −7.0 −8.0 ) K M ( φ 0 1 g o l −9.0 −12.0 −11.0 −10.0 −8.0 −7.0 −6.0 −9.0 MK Figure 6. Luminosity function in the first three iterations. clearly seen in the W92 model and the TMGS in Hammer- sley et al (1999)). This discrepancy could arise from their not having taken into account that the brightest stars in the bulge are up to 2 mag fainter than the disc giants (Frogel & Whitford 1987). This would shift the W92 luminosity func- tion to the right in Figure 7. It should be remembered that the W92 model was developed to predict the IRAS source counts. IRAS could see only the very top end of the bulge luminosity function, and the sources responsible are all dust- shrouded AGB stars. The dust enormously brightens the 12 and 25 micron fluxes over the expected photospheric flux. In fact, at the distance of the bulge, IRAS could not see purely photospheric stars at all. The TMGS, however, can detect normal bulge M giants (Frogel & Whitford 1987), not only AGBs, and the presence of dust leads only to a mi- nor increase in the K brightness. Therefore, in the TMGS while it is true that we do see the extreme AGB stars de- tected by IRAS, they in fact represent only a tiny fraction of the detected sources in each magnitude bin. Hence, the top end of the IRAS luminosity function and the top end of the TMGS luminosity function are dominated by different types of sources and so W92 could be close for IRAS but not get the top end of the K star counts correct. Between MK = −8 mag and MK = −6 mag (corre- sponding to the fainter limit of the TMGS at the distance of the bulge) the luminosity function of W92 does coincide with that determined here. As has already been noted, the result from the first iteration of the luminosity function (when the assumed density function was that of W92) is very close to the final result, particularly for absolute magnitudes fainter than MK < −8 mag. This implies that for the lines of sight used here the W92 model does correctly predict the num- ber of bulge stars per magnitude per square degree for −8 mag < MK < −6 mag, even though this model was aimed at matching the IRAS source counts. Given the match over this magnitude range we have chosen to use the W92 lumi- nosity function for the magnitudes fainter than MK = −6 mag, so that the luminosity function can be normalized. Comparison with the bolometric luminosity function obtained by other authors (see references in the introduc- Table 3. K-band luminosity function for bulge stars. MK (mag) log10 φ MK (mag) log10 φ –11.4 –11.2 –11.0 –10.8 –10.6 –10.4 –10.2 –10.0 –9.8 –9.6 –9.4 –9.2 –9.0 –8.50±0.50 –7.87±0.48 –7.94±0.43 –7.61±0.43 –7.36±0.44 –7.67±0.66 –7.83±0.83 –7.43±0.68 –7.33±0.79 –8.03±1.22 –7.45±0.85 –6.98±0.58 –6.60±0.46 –8.8 –8.6 –8.4 –8.2 –8.0 –7.8 –7.6 –7.4 –7.2 –7.0 –6.8 –6.6 –6.4 –6.35±0.33 –6.20±0.31 –6.19±0.28 –5.88±0.30 –5.50±0.20 –5.32±0.17 –5.30±0.22 –5.28±0.23 –5.10±0.17 –4.96±0.12 –4.87±0.19 –4.76±0.10 –4.63±0.14 tion) is not possible since bolometric corrections are not available. Also, in most of cases the magnitude interval is different. Tiede et al. (1995) provide, by combining data from different works, the luminosity function in the K band as a function of the apparent magnitude in the range 5.5 mag < mK < 16.5 mag. The brightest magnitudes are taken from Frogel & Whitford (1987). The comparison with our lumi- nosity function is not direct since they have not normalized their luminosity function to unity; moreover, they have not taken into account the narrow but non-negligible dispersion of distances. In Fig. 16 of Tiede et al. (1995) there is a fall- off in the luminosity function for mK ≤ 6.5 mag or in Fig. 18 of Frogel & Whitford (1987) for Mbol ≤ −4.2 mag, which could be comparable with that of our luminosity function at MK ≈ −8.0 mag. However, because of the much larger area covered by the TMGS, the error for the brightest magni- tudes is far lower in this paper, the result being pushed well above the noise; this is not the case for Frogel & Whitford (1987). The presented luminosity function for very bright stars (brighter than MK ∼ −9.5 mag) is of low precision. The number of bulge stars in this range is very small, so even small errors due to contamination from the spiral arms will mean that the luminosity function is overestimated and so the values should be taken as an upper limit. 5.1 Age of the bulge The age of the bulge is an open topic. There are authors who think the bulge is older than the halo (Lee 1992) whilst others hold the opposite opinion (Rich 1993). Although from the work presented here an accurate value for its age can- not be determined, the bulge is clearly older than the disc. The lack of very luminous stars in the bulge means that there are few supergiants and bright giants, and hence star formation regions. A comparison between the K-band lumi- nosity function derived here and models of stellar evolution could provide some further clue in this controversial subject. The model of Bertelli et al. (1994), with a 10-Gyr popula- tion and solar metallicity, predicts that all the stars should be fainter than MK = −8 mag, while these data show that there are some sources of −9.5 and −8 mag. This may indi- c(cid:13) 0000 RAS, MNRAS 000, 000–000 Inversion ... Bulge 11 l=5.4o, b=3.7o Wainscoat et al. 1992 − bulge Our data Eaton et al. 1984 − disk −4.0 −5.0 −6.0 −7.0 −8.0 0.80 0.60 0.40 0.20 ) 3 c p / s r a t s ( ) r ( D ) K M ( φ 0 1 g o l −9.0 −12.0 −10.0 −8.0 −6.0 MK 0.00 0.0 5000.0 10000.0 15000.0 r (pc) Figure 7. Luminosity function in the K-band (solid line). Com- parisons with W92 in the bulge and Eaton et al. (1984) in the disc are also provided. Figure 8. Density along the line of sight in the region (l = 5.4◦, b = 3.7◦). cate a mixture of populations with different ages embedded in the bulge. 6 DENSITY DISTRIBUTION 6.1 Density along the line of sight The second result is the density D(r) for each region (l, b), i.e. some points of the function D(~r) = D(r, l, b). ∆ is ob- tained by inversion of eq. (6) and then changing the variable in eq. (3) to recover D(r). As an example, the density distribution along the line of sight for one region (l = 5.4◦, b = 3.7◦) is shown in Fig. 8 after extinction correction. As can be seen, the bulge distribution of stars has a maximum around 8 kpc. There is a rise from ∼ 5 kpc to ∼ 8 kpc, and a fall off after this. Similar results were obtained in the other regions, except for some fluctuations due to errors (the errors in the counts may provide this fluctuation; see §4). The 37 regions used were the least noisy and least affected by patchy extinction. 6.2 Bulge cuts As was said in §2, the regions come from strips with constant declination: δ = −30◦, δ = −22◦ and δ = −16◦. The 37 regions used for density inversion are come from the strips at δ = −30◦, δ = −22◦ (as the bulge source density by δ = −16◦ is low). A strip can be thought of as a surface in space (Fig. 9) one axis is in R.A. (i.e. constant declination) and the other is distance along the line of sight, which can be converted to a distance parallel to the Sun–Galactic center line. Figures 10 and 11 show these plots with the z-axis representing the density. Note that the density scale (height) is different in both figures. As can be seen, there are two peaks and a valley in both figures. The valley only indicates the absence of data due to the fact that the Galactic plane between b = −2◦ and b = 2◦ c(cid:13) 0000 RAS, MNRAS 000, 000–000 δ=−30o δ=−22o 10.0 8.0 6.0 4.0 2.0 0.0 −2.0 −4.0 −6.0 −8.0 ) o ( b −10.0 −10.0 0.0 l (o) 10.0 Figure 9. Two constant-declination strips that cut the disc. The striped region is the Galactic plane zone, which was excluded. was avoided. If the plane data were included there would be only one peak. Galactic longitude increases and latitude decreases with increasing x. In Fig. 11, the left side (negative x) of the valley has a lower density than the right side (positive x) due to the abrupt fall-off of the density with distance from the Galactic center. This is not observed in Fig. 10 because this strip almost cuts across the Galactic center so both sides of the valley are nearly symmetric. When comparing the position of the peaks, and hence the maximum density, in both figures, the peaks are notice- ably closer to the Sun for δ = −22◦ (l = 7.5◦) than for δ = −30◦ (l = −1◦). The non-axisymmetry of the bulge is the most plausible explanation for this and the bulge is closer to us at higher galactic longitudes. This can, in fact, be seen in the individual strips, as the left peak (i.e. larger l) is closer than the right one in both figures. Hammersley et 12 L´opez-Corredoira et al. 1.60 12000 y (pc) 8000 3 ) c p / s r a t s ( y t i s n e D 4000 0 -2000 0 x (pc) 2000 Figure 10. Plot of the density (height) as a function of both spatial coordinates defined by a cut of the bulge in δ = −30◦. Galactic latitude is increased from left to right (x-axis). The y- axis is distance parallel to the line joining the Sun to the Galactic Centre. The grid scale is 400 pc for each small square. The range of distances is from 4000 to 12000 pc along the line of sight, and from −2000 to +2000 pc in the x-axis. The origin is at the Sun. the previous subsection argued for non-axisymmetry in the bulge, so the next stage was to determine the parameters. Ellipsoids were used for the fit, with two axes in the Galactic plane and a third perpendicular to these. The pos- sible tilt of the bulge out of the plane was neglected as there is no evidence for this (Weiland et al. 1994). Also the posi- tion of the Sun 15 pc above the plane (Hammersley et al. 1995) does not have a significant influence since the bulge extends much further from the plane. The Galactocentric distance along the major axis for different isodensity ellipsoids is 1 + K 2 2 x2 2 + K 2 z z2 t =px2 (22) and the distance along the minor axis is t/Kz. The pro- jections of the vector distance to the Galactic centre are represented x1 and x2 (Fig. 12), and z is the distance to the plane. K2 and Kz are the axial ratios between axes x1 and x2, and x1 and z, respectively. Both ratio are defined to be greater than one. From the same figure x1 and x2 are defined as follows: x1 = R cos(β − α) and x2 = R sin(β − α), with R =p(r cos b)2 + R2 z = r sin b, 0 − 2rR0 cos b cos l, and, following the sine rule, β = sin−1 r cos b sin l . R (23) (24) (25) (26) (27) 0.47 3 ) c p / s r a t s ( y t i s n e D 12000 y (pc) 8000 4000 0 -2000 0 x (pc) 2000 Figure 11. The same plot as Figure 10 but for δ = −22◦. al. (1999: Sect. 7, Fig. B) also show this asymmetry derived from TMGS data. 6.3 The three-dimensional bulge The morphology of the bulge can be examined by fitting the isodensity surfaces to D(~r) = D(r, l, b). The results of The ellipsoids have four free parameters: R0, the Sun- Galactic centre distance (the ellipsoids are then centred on this position); Kz and Ky, the axial ratios with respect to the major axis (x); and α, the angle between the major axis of the triaxial bulge and the line of sight to the Galactic centre (α between 0◦ and 90◦ is where the tip of the major axis lies in the first quadrant). Three-dimensional ellipsoids are fitted to 20 isodensity surfaces (from 0.1 to 2.0 star pc−3, in steps of 0.1) with the four free parameters. The four parameters are then averaged for the 20 ellip- soids and the results are: R0 = 7860 ± 90 pc, K2 = 1.87 ± 0.18, Kz = 3.0 ± 0.9 and α = 12 ± 6 deg. (28) The errors are calculated from the average of the ellip- soids and so do not include possible systematic errors (for example: subtraction of the disc, contamination from other components, methodological inaccuracies of the inversion, etc.), which are difficult to determine. However, by far the largest effect on the bulge counts is the massive asymme- try in the counts caused by the triaxiality of the bulge, as c(cid:13) 0000 RAS, MNRAS 000, 000–000 Inversion ... Bulge 13 C β R α x 1 . S R 0 l r cos(b) x 2 P Figure 12. Cut of an ellipsoidal bulge in the Galactic plane. C is the Galactic centre, P is a given point on the ellipsoid and S is the Sun. shown in Hammersley et al. (1999). The other systematic ef- fects are at least an order of magnitude below this, so while they do have an effect, it is small. Hence, the true errors are larger than stated but tests suggest that they do not alter the general findings presented here. The error in Kz is quite large and is due to the non- constant axial ratio of the ellipsoids. Kz tends to increase towards the centre, i.e. the outer bulge is more circular than the inner bulge. This will be further discussed in §6.4. 6.3.1 Axial ratios and orientation The axial ratios of the bulge are 1:0.54:0.33. These numbers indicate that the bulge is triaxial with the major axis close to the line of sight towards the Galactic centre. In general, the result presented here are in agreement with those from other authors. The projection, as viewed from the position of the Sun, of an ellipsoid of the above characteristics, gives an ellipse with axial ratio 1.7± 0.5 (i.e. 1:0.58). This is com- patible with the value of 1:0.6 obtained by Weiland et al. (1994) or 1:0.61 by (Kent et al. 1991). From a dynamic model assuming a gas ring in a steady state, Vietri (1986) finds axial ratios of 1:0.7:0.4, which is close to our result. Binney et al. (1991) found α = 16 deg for a bar, i.e. a triaxial structure in the centre of the Galaxy, in order to explain the kinematics of the gas in the centre of the Galaxy. Weinberg (1992) gives α = 36 ± 10 deg and K2 = 1.67 from his analysis of IRAS data. More recently, Nikolaev & Weinberg (1997) obtained a bar from IRAS sources with α = 19 deg and K2 between 2.2 and 2.7. Stanek et al. (1997), based on the analysis of optical photometric data for regions of low extinction, predicted an α between 20 and 30 deg and 1:0.43:0.29 axial ratios, which is also quite close to the result presented here. Various authors have examined the COBE- DIRBE flux maps for triaxiality: Dwek et al. (1995) give higher eccentricity values for the axial ratios, 1:0.33:0.22, c(cid:13) 0000 RAS, MNRAS 000, 000–000 but the angle α = 20 ± 10 deg is compatible with the value given here; Binney et al. (1997) derive 1:0.6:0.4 ratios and an angle α ∼ 20 deg; Freudenreich (1998) obtained a best fit with 1:0.38:0.26 ratios and α = 14 deg. Normally, when the low latitudes are excluded the fit of the triaxial bulge has an angle of about 25 degrees (Sevenster et al. 1999). The majority of the above authors did not use inversion; rather they fitted the flux or the star counts to models using a priori assumptions. Binney et al. (1997) were the exception in using inversion on the COBE-DIRBE surface-brightness mapsk and, on the basis of specific assumption, obtained results close to those presented here. 6.3.2 Galactocentric distance The distance R0 derived here is slightly less than that used in the W92 model of the disc (8.5 kpc). However, the small changes in R0 can be compensated by small changes in the other model parameters, such as the scale length, so that the predicted counts remain the same. As the model used already gave a good fit to the disc, we decided not to make ad hoc modifications to account for a smaller R0 since the disc is not the subject in this paper. The lack of previous assumptions makes the determi- nation of R0 presented here different from those of other authors. In particular, no information is required on the ob- jects observed. However, the values determined here are very close to the currently accepted value of just under 8 kpc. Reid et al. (1988) deduce a value R0 = 7.1 ± 1.5 kpc from k The inversion of the flux and the inversion of the star counts are significantly different. Since star counts provide a function for each region of space and the flux is only one number for each of those regions, the inversion of the flux is less suitable for directly extracting information from the data and further assumptions are needed. 14 L´opez-Corredoira et al. Table 4. Relationship between the maximum distance of the el- lipsoid and the bulge star density. t (pc) D (pc−3) t (pc) D (pc−3) 3020 ± 810 2630 ± 650 2420 ± 590 2230 ± 490 2120 ± 450 1990 ± 380 1900 ± 350 1840 ± 330 1720 ± 280 1670 ± 270 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1620 ± 250 1580 ± 240 1540 ± 240 1460 ± 230 1420 ± 230 1390 ± 230 1380 ± 240 1360 ± 250 1360 ± 260 1320 ± 220 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 1e+01 1e+00 1e−01 ) 3 c p / s r a t s ( y t i s n e D direct observations of Sgr B2. Gwinn et al. (1992), by means of observations of masers in W49, derive R0 = 8.1 ± 1.1 kpc. Moran (1993) obtains R0 = 7.7 kpc, from OH/IR stars distances. Turbide & Moffat (1993) obtain R0 = 7.9 ± 1.0 kpc, from measurements of the distances to young stars by means of CCD photometry and assuming that there is no metallicity gradient in the outer regions of the Galaxy; al- though they get 7.2 kpc when a certain gradient is assumed. Paczy´nski & Stanek (1998) derived R0 = 7.97 ± 0.08 (sys- tematic effects make the true error larger) from the compar- ison between Hipparcos and OGLE data. Olling & Merrifield (1998a, 1998b) obtain R0 = 7.1 ± 0.4. Etc. Generally, many studies based on indirect measurements claim the Galacto- centric distance to be somewhat less than 8.0 kpc (see also the review by Reid, 1993). 6.3.3 Density as a function of the distance to the Galactic centre A power law with exponent −1.8 is observed in the cen- tre of the bulge and also in other galaxies (Becklin & Neugebauer 1968; Sanders & Lowinger 1972; Maihara et al. 1978; Bailey 1980; see review by Sellwood & Sanders 1988). When the density function D(t) (Table 4) is fitted to D(t) = A(t/t0)1.8 exp(−(t/t0)γ ), with γ, t0 and A as free parameters, then we obtain D(t) = 1.17(t/2180 pc)−1.8 exp(−(t/2180 pc)1.8) star pc−3. (29) This gives an estimate of the fall-off in density between 1.3 and 3.0 kpc from the centre in the direction parallel to the major axis or between 0.4 and 1.0 kpc in the direction perpendicular to the plane. As can be seen in Fig. 13, the dispersion of points around this law is large, so it is possible to accommodate other functions or even a different set of parameter. A different luminosity function amplitude would change the amplitude of the stellar density, A. If the nor- malization for the luminosity function were incorrect then the factor needed to multiply the luminosity function would be used to divide the star-density amplitude. 0.0 1000.0 2000.0 t (pc) 3000.0 4000.0 Figure 13. Fit of the density distribution. The solid line is the best fit using eq. (29). Figure 14. N (mK = 9.0 mag) along the three strips that are used with constant declinations: δ = −30◦, which cuts the plane at l = −1◦; δ = −22◦, which cuts the plane at l = 7◦; and δ = −16◦, which cuts the plane at l = 15◦ once the W92 disc and bulge (according to eqs. (28) and (29)) are subtracted. 6.3.4 Goodness of the inversion The residual counts for mK < 9 after subtracting both the bulge determined here and the W92 disc model from the original counts are plotted in Figure 14. As can be seen, the off-plane residual counts (the b < 2◦ regions are clearly contaminated by other components) are reduced to typically a few per cent of the original counts shown in Figure 2. For the δ = −30◦ strip the residuals are typically 100 star/deg2 compared to the 1500 star/deg2 in the original counts. Hence the proposed bulge parameters do accurately reproduce the observed counts. c(cid:13) 0000 RAS, MNRAS 000, 000–000 6.3.5 A triaxial bulge From Figs. 10 and 11, the non-axisymmetry was determined for the plane. Furthermore, the axial ratio K2 is close to 2 (and not 1, the condition of axisymmetry). Therefore the bulge is a triaxial ellipsoid orientated in such a way that the minor axis is perpendicular to the Galactic plane, and the angle between the major axis and the Sun–Galactic centre line is 12◦ in the first quadrant. Whether this structure is called a bar or triaxial bulge is not only a question of wording. Apart from the morphol- ogy, the population is also has to take into account: bulges are older than bars (Kuijken 1996), though both are older than the disc. Precise calculations of the age (see §5.1) would be necessary to differentiate between them. However, there is evidence of another lengthened structure, a bar, (Ham- mersley et al. 1994; Calbet et al. 1996; Garz´on et al. 1997) whose angle is ∼ 75◦ in the first quadrant. This has major star formation regions at both extremes (towards l = 27◦ and l = −22◦) and there is evidence for a preceding dust lane (Calbet et al. 1996). If this other component exists then the structure discussed in this paper must be called a "bulge", unless we are prepared to entertain the notion that the Galaxy has two bars. 6.4 Bulge with variable Kz ellipsoids A large error in Kz is obtained when it is assumed constant, as was indicated in the previous subsection. Therefore, it is possible that the Kz values are not constant, and so another dependence on the isodensity contours was tried. When the ellipsoids are fitted allowing a linearly variable Kz, then (30) Kz = (1.66 ± 0.17) + (1.73 ± 0.14)D (where the units of D are star pc−3), whose weighted average is Kz = 3.0, as obtained in eq. (28). The other parameters (K2, R0 and α) remain nearly constant with respect to D. This variation of Kz is independent of the trial solution in the iteration process (see §4.1). A fourth iteration was performed for both the luminosity function and the density with the feed-back of the variable Kz, and it could be seen that the same parameters are recovered again, within a 1-σ error. Indeed, the x1–z ratio is Kz = (1.76 ± 0.32) + (1.70 ± 0.27)D. (31) This linear dependence is valid in the density interval from 0.1 to 2.0 star pc−3. For highest densities, Kz is ex- pected to grow more slowly. Kz can also be expressed as a function of t, although this dependence is non-linear. The fit to an exponential law is: Kz = (8.4 ± 1.7) exp(cid:18) −t (2000 ± 920) pc(cid:19) (32) and is valid for the range of distances, t, used here (see Fig. 16). Figure 15 shows the variation of eccentricity as a function of the density. With Kz so defined, the density, D, is given by Table 5 or Figure 16.⋆⋆. Inversion ... Bulge 15 3000.0 2000.0 1000.0 0.0 −1000.0 −2000.0 ) c p ( z −3000.0 −3000.0 −2000.0 −1000.0 0.0 x1 (pc) 1000.0 2000.0 3000.0 Figure 15. Cut of the bulge in the x1–z plane when Kz is given by eq. (32). The ellipses represent isodensity lines between 0.1 and 2.0 star pc−3, with intervals of 0.1. Table 5. Relationship between the distance along the minor axis, t/Kz and the density, when Kz is given by (32). t/Kz (pc) D (star pc−3) t/Kz (pc) D (star pc−3) 1400 ± 380 1170 ± 290 1010 ± 240 880 ± 200 780 ± 160 700 ± 140 640 ± 120 570 ± 100 540 ± 90 510 ± 80 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 470 ± 60 440 ± 50 430 ± 40 410 ± 40 390 ± 40 380 ± 30 370 ± 30 360 ± 30 350 ± 30 330 ± 20 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 The best fit to a law of type D(t/Kz) = A(t/(Kzt0))−1.8 exp(−(t/(Kz ×t0))γ ) is: D(t/Kz) = 0.106(cid:18) t/Kz 1820 pc(cid:19)−1.8 exp −(cid:18) t/Kz 1820 pc(cid:19)5.4! star pc−3. (33) 6.4.1 Goodness of the inversion The residual counts after both the bulge determined here with variable Kz and the W92 disc model are subtracted from the original counts for mK < 9 are plotted in Figure 17. As can be seen the residuals are now somewhat lower than when Kz is constant (Fig. 14). Typically the residuals are now 50 to 100 star/deg2 and the maximum has fallen from 300 star/deg2 with constant Kz to 200 star/deg2. Therefore, ⋆⋆ The density is expressed as a function of t/Kz, the distance along the z-axis, because the variation of Kz with t fluctuates too much. The ellipsoid size decreases when the density, D, increases; however, Kz increases with D, so the axis x1 increases. Hence, the variation of D as a function of t is too sensitive to noise. c(cid:13) 0000 RAS, MNRAS 000, 000–000 16 L´opez-Corredoira et al. 1e+01 1e+00 1e−01 1e−02 ) 3 c p / s r a t s ( y t i s n e D 1e−03 0.0 500.0 1000.0 t/Kz (pc) 1500.0 2000.0 Figure 16. Fit of the density distribution when Kz is given by eq. (32). The solid line stands for the fit to (33). Sun direction Figure 18. Projection of the bulge, eq. (32), when it is observed face-on (integration of z direction). The square is 8 kpc×8 kpc centred on the Galactic Centre. The outer contour represents 10 star pc−2, the second contour stands for 410 star pc−2, etc., and the inner contour stands for 4810 star pc−2 (the interval between consecutive contours is 400 star pc−2). superposition of two components is being observed, i.e the bulge and another structure, a bar, closer to the plane. If this were true, the luminosity function would have two different populations, especially in the regions closest to the plane. A gradient within the bulge is also possible. A greater number of bright stars in the innermost bulge (as observed by Calbet et al. 1995), with a smooth variation from the inner to the outer bulge, could be responsible of this effect. Both of these causes would lead to a gradient in the luminosity function. However, as the luminosity function has been assumed to be constant the result after inversion would be a gradient in Kz. Tests on the data indicate that this is possible. Giving the luminosity function a gradient in z, but such that the luminosity function remains within the error bars for a de- termined average function, is sufficient to produce changes in the observed gradient in Kz. In any case, the errors in the luminosity function (see Table 3) do limit this variation of populations. 6.5 Stellar content of the bulge Integrating the density over all space will give the stellar content of the whole bulge: N =Z D(t)dV . (34) The volume element dV , under a change of variable to elliptic coordinates t, θ and φ, is related to the Cartesian coordinates x = t sin θ cos φ, y = t cos θ, K2 through sin θ sin φ, z = t Kz Figure 17. N (mK = 9.0 mag) along the three strips that are used with constant declinations: δ = −30◦, which cuts the plane at l = −1◦; δ = −22◦, which cuts the plane at l = 7◦; and δ = −16◦, which cuts the plane at l = 15◦ once the W92 disc and bulge (according to eqs. (28), (32) and (33)) are subtracted. the variable Kz does provide a better fit to the observed counts. The aspect of the bulge as seen by an observer far away in the z-axis, i.e. the Galaxy observed face-on, would be as shown in Figure 18. The sharp fall-off in density is very noticeable. The bulge in a face-on Milky Way-like Galaxy presents, according to our results, a very high contrast be- tween central regions (with up to 10000 star pc−2) and re- gions at 3 kpc in the major axis (with 100 star pc−2). Whether this variation of Kz is a true feature of the density distribution or not is a matter for further investiga- tion. However, this is observed in other galaxies (Varela et al. 1993) and we do not believe that the result of this sub- section is due to systematic errors, although this possibility cannot be totally excluded. Other possible causes for this might be either that a dV = hthθhφdtdθdφ, (35) c(cid:13) 0000 RAS, MNRAS 000, 000–000 with hi =s(cid:18) ∂x ∂qi(cid:19)2 ∂qi(cid:19)2 +(cid:18) ∂y ∂qi(cid:19)2 +(cid:18) ∂z . Hence, (36) cos2φ K 2 2 dV = t2 sin θrsin2 φ + ×s sin2 θ ×s cos2 θ + cos2 θ(cid:20)cos2 φ + + sin2 θ(cid:20)cos2 φ + K 2 z K 2 z sin2 φ K 2 2 (cid:21) 2 (cid:21)dtdθdφ. sin2 φ K 2 The result is 2.8 × 1010 stars for Kz = 3 with D from eq. (29); and 4.1 × 1010 stars with a variable Kz from eq. (32) and D from eq. (33), i.e. a factor 1.4 greater. This is, of course, only an estimation which includes an extrapolation of D to all space and the assumption of a correct luminosity- function normalization (see §5). Nevertheless, it leads to an order of magnitude for the mass of the Galactic bulge (taking an average mass for a star of ∼ 1M⊙) compatible with other data (for instance, ∼ 2 × 1010 M⊙ in Gould 1997); so this supports the normalization and the extrapolation. From the integration of the luminosity function it is found that TMGS stars from the whole bulge (mK < 9.0 mag) represent only a fraction (∼ 2×10−5) of the total number of stars, i.e. 6×105 stars for Kz = 3. 7 HOW DIFFERENT WOULD THE RESULTS FOR A DIFFERENT DISC MODEL BE? Errors in different parts of the inversion procedure used here will lead to changes in the results. One important source of error may be the disc model that is used: were a different model to be used, the answers would be different. Clearly the answer to a certain extent depends on the new model to use. As has been detailed earlier in §3.3, the disc model used here is in good agreement with the observed TMGS star counts where the disc is isolated and so the expectation is that its extrapolation to regions where bulge and disc are observed will lead only to small errors in the counts. By definition, the bulge is an excess over the extrapolated disc in central regions of the Galaxy so, also by definition, the error of the present disc model cannot be very large once its fitting to observational data in external parts of the Galaxy has been tested. From the integral equation (6), it can be deduced that these errors, δNK,disc(mK), follow for all regions (l, b): δNK,disc(mK ) = −ω δΦK,bulge(mK + 5 − 5 log10 ρK)∆bulge,K(ρK)ρ2 K dρK 0 ×Z ∞ −ωZ ∞ × ρ2 0 KdρK, ΦK,bulge(mK + 5 − 5 log10 ρK)δ∆bulge,K (ρK) (38) c(cid:13) 0000 RAS, MNRAS 000, 000–000 (37) δDdisc = Ddisc δhR(R − R⊙) h2 R , (39) Inversion ... Bulge 17 If we know δNK,disc(mK), which differentiates the "real disc" from our model, we could derive how large δΦ and δ∆ are. The inversion procedure explained in §4 produces solutions which are close when we begin the iteration from counts that are similar, as can be seen from eq. (13). Hence, for small δNK,disc(mK), δΦ and δ∆ are also small. That is, the behaviour is not a chaotic such that small departures from the original counts would not produce very different solutions. For instance, let us suppose that there is an error, δhR, in the scale length of the disc (equal to 3.5 kpc in the W92 model we assumed). This leads to an error in the density due to the disc of where R is distance from the Galactic centre and R⊙ is this distance for the Sun (8.5 kpc in the W92 model). Hence, by means of eqs. (3), (4) and (5) for the disc, 0 (40) K dρK, δhR h2 δNK,disc(mK) = ω ΦK,disc(mK + 5 − 5 log10 ρK) R Z ∞ × ∆disc,K (ρK)(R − R⊙)ρ2 which can be set equal to expression (38) for all mK , l and b. However, whilst in principle the change in the disc is pro- portional to the scale length, and there are certainly values quoted in the literature as low as 2.2 kpc (Ruphy et al. 1996), it should be remembered that it is already known that the W92 model gives an excellent fit in the areas where the disc is isolated. Therefore, were one to alter the scale length, then other parameters also would have to be varied to compensate, otherwise the excellent agreement would be lost. It would be difficult to change the disc more than a few per cent without the effect becoming noticeable. In the selected regions the bulge counts are dominant. For instance, the maximum contribution of the disc in the region (l = 0.3◦, b = −2.0◦) is 1200 star/deg2 up to 9th K-magnitude whereas the total counts are around 6000 star/deg2 (Fig. 2), i.e. in this case only 20% of the sources are from the disc. The ratio varies according to the region examined, but in most of the regions used the bulge is the dominant feature. Furthermore the error in the number of bulge sources is determined by the error in the number of disc sources, therefore if the relative proportion of the disc sources is low and the disc model gives a good fit to the TMGS counts this implies that the error introduced to the bulge counts will be of the order of a few per cent, probably below the Poissonian noise. Therefore, the errors in the disc affects the shape and luminosity function of the bulge only slightly. 7.1 Experiments of inversion varying the parameters of the disc or the extinction A simple test can be carried out to verify what has been claimed in this section: small changes in the parameters of the disc (or also the extinction) do not greatly affect the results, i.e. there is no chaotic behaviour. We run the same inversion programs again to obtain both the luminosity function and the density distribution. Two examples are shown in this subsection: a) inversion with 18 L´opez-Corredoira et al. −4.0 −5.0 ) Κ Μ ( φ 0 1 g o l −6.0 −7.0 −8.0 −9.0 −12.0 hR=3.5 kpc, AK=0.07 mag kpc−1 (reference) hR=3.0 kpc, AK=0.07 mag kpc−1 hR=3.5 kpc, AK=0.05 mag kpc−1 W92 − bulge −10.0 −8.0 −6.0 MK Figure 19. Luminosity function with different parameters for the disc model and the extinction as well as for the W92 model of the bulge. Reference luminosity function is the one obtained in §5. hR = 3.0 kpc instead of the original value of hR = 3.5 kpc; b) inversion with extinction normalization coefficient AK = 0.05 mag kpc−1 instead of AK = 0.07 mag kpc−1. The new luminosity functions are shown in Fig. 19 in comparison with that obtained in section 5. Both inversions with new disc and extinction are fit to constant axial-ratio triaxial ellipsoids respectively with parameters: a) R0 = 8400 ± 190 pc, Kz = 2.5 ± 1.3, Ky = 1.75 ± 0.05, α = 12◦ ± 3◦; b) R0 = 7600 ± 130 pc, Kz = 4.1 ± 1.1, Ky = 1.70 ± 0.05, α = 9◦ ± 2◦. These values are close to those obtained in §6.3, which is an indication of the robustness of the method of inversion. In case a), the luminosity function for very bright stars in K is higher than the reference one in comparison with the faintest parts due, perhaps, to a defect of outer bulge stars. The disc model in a) is unrealistic and provides fur- ther star counts in the Galactic centre than there should be (∼ 25% more stars than in the reference model); the outer regions of the bulge would have zero, or negative, counts once the disc is subtracted, so they do not contribute to the weighted average of the luminosity function. In case b), the Galactic centre is closer to us, as expected if the extinction is lower. No physical meanings can be derived from these ex- periments, since the disc model in a) or the extinction model in b) is less exact than that in the reference case. They sim- ply provide a verification of the robustness of the inversion method. 8 CONCLUSIONS The procedure used here is rather different from that of those authors who fit the parameters directly to the star counts. First, the counts were inverted. Then, after the lumi- nosity function and density distribution were evaluated and an approximate ellipsoidal shape was evident, the parame- ters could be fitted for each isodensity surface. Assuming an ellipsoidal bulge with constant parameters for all isodensity regions and fitting these parameters to the counts is less rig- orous since there is no a priori evidence for this assumption. In fact, our method suggests that constant parameters for the ellipsoids do not give the best fit for the density, D(~r). Instead, a decreasing major–minor axial ratio from inside to outside would provide best results. These results are: The distance to the centre of the bulge, i.e. the centre of the Galaxy, is 7.86 ± 0.09 kpc (systematic effects make the true error larger). The relative abundance of the brightest sources in the bulge (MK < −8.0 mag) is much less than in the disc. The bulge is triaxial with axial ratios 1:0.54:0.33, the mi- nor axis perpendicular to the Galactic plane, and the major axis nearly along the line of sight to the Galactic centre. The best fit giving an angle equal to 12 deg shifted to positive Galactic longitudes in the plane in the first quadrant. A gradient in the major–minor axial ratio is measured. However, there are various possible caused which include eccentricity of the true density-ellipsoid gradient or a popu- lation gradient. The stellar density drops quickly with distance from the Galactic centre (i.e. the density distribution is sharply peaked). The −1.8 power-law observed at the Galactic cen- tre needs to be multiplied by an exponential to account for the fast drop in density in the outer bulge. REFERENCES Bahcall J. N., 1986, ARA&A 24, 577 Bahcall J. N., Soneira R. M., 1980, ApJS 44, 73 Bailey M. E., 1980, MNRAS 190, 217 Bal´azs L. G., 1995, Inverse Problems 11, 731 Becklin E. E., Neugebauer G., 1968, ApJ 151, 145 Bertelli G., Bressan A., Chiosi C., Fagotto F., Nasi E., 1994, A&AS 106, 275 Binney J., Gerhard O. E., Stark A. A., Bally J., Uchida K. I., 1991, MNRAS 252, 210 Binney J., Gerhard O., Spergel D., 1997, MNRAS 288, 365 Blitz L., Spergel D. N., 1991, ApJ 370, 205 Bok B. J., 1937, The Distribution of Stars in Space, University of Chicago Press, Chicago Buser R., Kaeser U., 1983, in: The Nearby Stars and the Stellar Luminosity Function, IAU Symp. 76, ed. A. G. Davis Phillip, A. R. Upgren, Schenectady, New York. p. 147 Calbet X., Mahoney T., Garz´on F., Hammersley P. L., 1995, MN- RAS 276, 301 Calbet X., Mahoney T., Hammersley P. L., Garz´on F., L´opez- Corredoira M., 1996, ApJ 457, L27 Cohen M., 1994a, AJ 107(2), 582 Cohen M., 1994b, Ap&SS 217(1), 181 Craig I. J. D., Brown J. C., 1986, Inverse problems in astronomy, Adam Hilger, Bristol-Boston Davidge T. J., 1991, ApJ 380, 116 De Poy D. L., Terndrup D. M., Frogel J. A., Atwood B., Blum R., 1993, AJ 105, 2121 Dwek E., Arendt R. G., Hauser M. G., et al., 1995, ApJ 445, 716 Eaton N., Adams D. J, Gilels A. B., 1984, MNRAS 208, 241 Feast M. W. & Whitelock P. A., 1990, in: in: Bulges of galaxies, B. J. Jarvis, D. M. Terndrup, eds., Garching: ESO, p. 3 Freudenreich H. T., 1998, ApJ 492, 495 Frogel J. A., 1988, ARA&A 26, 51 Frogel J. A., 1990, in: Bulges of galaxies, B. J. Jarvis, D. M. Terndrup, eds., Garching: ESO, p. 111 c(cid:13) 0000 RAS, MNRAS 000, 000–000 Inversion ... Bulge 19 spective, S. R. Majewski, ed., ASP Conference Series, Vol. 49, San Francisco, p. 65 Robin A. C., Cr´ez´e M., 1986, A&A 157, 71 Ruelas-Mayorga R. A., 1991, Rev. Mex. Astron. Astrof. 22, 27 Ruelas-Mayorga A., Noriega-Mendoza H., 1995, Rev. Mex. As- tron. Astrof. 31, 115 Ruphy S., Robin A. C., Epchtein N., et al., 1996, A&A 313, L21 Sanders R. H., Lowinger T, 1972, AJ 77, 292 Scoville J. Z., Young N. S., Lucy L. B., 1983, ApJ 270, 443 Sellwood, J. A., Sanders R. H., 1988, MNRAS 233, 611 Sevenster M. N., 1996, in: Barred galaxies, IAU symp. 157, Buta R., Crocker D. A., Elmegreen B. G., eds., ASP conference series, San Francisco., p. 536 Sevenster M., Prasenjit S., Valls-Gabaud D., Fux R., 1999, MN- RAS 307, 584 Stanek K. Z., Mateo M., Udalski A., et al., 1994, ApJ 429, L73 Stanek K. Z., Mateo M., Udalski A., et al., 1996, in: Barred galax- ies, IAU symp. 157, Buta R., Crocker D. A., Elmegreen B. G., eds., ASP conference series, San Francisco, p. 545 Stanek K. Z., Udalski A., Szyma´nski M., et al., 1997, ApJ 477, 163 Tiede G. P., Frogel J. A., Terndrup D. M., 1995, AJ 110(6), 2788 Tyson N. D., Rich R. M., 1993, in: Galactic Bulges, IAU Symp. 153, H. Dejonghe, H. J. Habing, eds., Kluwer, Dordrecht, p. 333 Trumpler R. J., Weaver H. F., 1953, Statistical Astronomy, Uni- versity California Press, Berkeley, ch. 5 Turbide L., Moffat F. J., 1993, AJ 105, 1831 Turchin V. F., Kozlov V. P., Malkevich M. S., 1971, Soviet Physics Uspekhi 13(6), 681 Varela A. M., Simonneau E., Munoz-Tun´on C., 1993, in: Galac- tic Bulges, IAU Symp. 153, H. Dejonghe, H. J. Habing, eds., Kluwer, Dordrecht, p. 435 Vietri M., 1986, ApJ 306, 48 Wainscoat R. J., Cohen M., Volk K., Walker H. J., Schwartz D. E., 1992, ApJS 83, 111 (W92) Weiland J. L., Arendt R. G., Berriman G. B., et al., 1994, ApJ 425(2), L81 Weinberg M. D., 1992, ApJ 384, 81 Whitelock P. A., Feast M. W., Catchpole R. M., 1991, MNRAS 248, 276 Wo´zniak P. R., Stanek K. Z., 1996, ApJ 464, 233 Frogel J. A., Terndrup D. M., Blanco V. M., Whitford A. E., 1990, ApJ 353, 494 Frogel J. A., Tiede G. P., Kuchinski L. E., 1999, AJ 117, 2296 Frogel J. A., Whitford A. E., 1987, ApJ 320, 199 Garz´on F., Hammersley P. L., Mahoney T., et al., 1993, MNRAS, 264, 773 Garz´on F., Hammersley P. L., Calbet X., Mahoney T. J., L´opez- Corredoira M., 1996, in: New Extragalactic Perspectives in the New South Africa, D. L. Block, J. Mayo Greenberg, eds., Kluwer, Dordrecht, p. 388 Garz´on F., L´opez-Corredoira M., Hammersley P. L., et al., 1997, ApJ 491, L31 Gerhard O. E., Binney J., Zhao H., 1998, in: Highlights of As- tronomy Vol. 11 (23rd. General Assembly of the IAU), J. An- dersen, ed., p. 628 Gilmore G., 1984, MNRAS 207, 223 Gilmore G., 1989, in: The Milky Way as Galaxy, R. Buser, I. King, eds., SAAS-FEE, Sauverny-Versoix, ch. 2 Gould A., 1997, Sheffield workshop on Identification of Dark Mat- ter, N. J. C. Spooner, ed., World Scientific, Singapore, p. 170 Gwinn C. R., Moran J. M., Reid M. J., 1992, ApJ 393, 149 Hammersley P. L., Cohen M., Mahoney T. J., Garz´on F., L´opez– Corredoira M., 1999, MNRAS, 308, 333 Hammersley P. L., Garz´on F., Mahoney T., Calbet X., 1994, MN- RAS 269, 753 Hammersley P. L., Garz´on F., Mahoney T., Calbet X., 1995, MN- RAS 273, 206 Holtzman J. A., Watson A. M., Baum W. A., et al., 1998, AJ 115, 1946 Houdashelt M. L., 1996, PASP 108, 828 Ibata R. A., Gilmore G. F., 1995, MNRAS 275, 605 Jupp D. L. B., Vozoff, 1975, Geophys. J. R. astr. Soc. 42, 957 Kent S. M., Dame T. M., Fazio G., 1991, ApJ 378, 131 Kuijken K., 1996, in: Unsolved problems of the Milky Way, IAU Symp. 169, L. Blitz, P. Teuben, eds., Kluwer, Dordrecht, p. 71 Lee Y. W., 1992, PASP 104, 798 L´opez-Corredoira M., Garz´on F., Mahoney T., Hammersley P., 1997a, in: The Impact of Large Scale Near-IR Sky Surveys, F. Garz´on, N. Epchtein, A. Omont, B. Burton, P. Persi, eds., Kluwer, Dordrecht, p. 107 L´opez-Corredoira M., Garz´on F., Hammersley P. L., Mahoney T. J., Calbet X., 1997b, MNRAS 292, L15 Loredo T. J., 1990, in: Maximum Entropy and Bayesian Methods, P. F. Foug`ere, ed., Kluwer, Dordrecht, p. 81 Lucy L. B., 1974, AJ 79(6), 745 Lucy L. B., 1994, A&A 289, 983 Maihara T., Oda N., Sugiyama T., Okuda H., 1978, PASJ 30, 1 Mihalas D., Binney J., 1981, in: Galactic Astronomy, WH Free- man, San Francisco Minniti D., 1996, ApJ 459, 175 Moran J. M., 1993, in: Sub Arcsecond Radio Astronomy, ed. R. J. Davis, R. S. Booth, Cambridge University Press, Cambridge, p. 62 Nakada Y., Deguchi S., Hashimoto O., et al., 1991, Nat 353, 140 Nikolaev S., Weinberg M. D., 1997, ApJ 487, 885 Olling R. P., Merrifield M. R., 1998a, in: Galactic halos: A UC Santa Cruz Workshop (ASP Conf., 136), D. Zaritsky, ed., p. 216 Olling R. P., Merrifield M. R., 1998b, MNRAS 297, 943 Ortiz R., L´epine J. R. D., 1993, A&A 279, 90 Paczy´nski B., Stanek K., 1998, ApJ 494, L219 Prichet C., 1983, AJ 84, 1476 Reid M. J., Schneps M. H., Moran J. M., et al., 1988, ApJ 330, 809 Reid M. J., 1993, ARA&A 31, 345 Rich R. M., 1993, in: Galactic Evolution: The Milky Way Per- c(cid:13) 0000 RAS, MNRAS 000, 000–000
astro-ph/0305397
1
0305
2003-05-21T12:49:04
Rotational periods of very young brown dwarfs and very low-mass stars in ChaI
[ "astro-ph" ]
We have studied the photometric variability of very young brown dwarfs and very low-mass stars (masses well below 0.2 M_sun) in the ChaI star forming region. We have determined photometric periods in the Gunn i and R band for the three M6.5-M7 type brown dwarf candidates ChaHa2, ChaHa3 and ChaHa6 of 2.2 to 3.4 days. These are the longest photometric periods found for any brown dwarf so far. If interpreted as rotationally induced they correspond to moderately fast rotational velocities, which is fully consistent with their v sini values and their relatively large radii. We have also determined periods for the two M5-M5.5 type very low-mass stars B34 and CHXR78C. In addition to the Gunn i and R band data, we have analysed JHK_s monitoring data of the targets, which have been taken a few weeks earlier and confirm the periods found in the optical data. Upper limits for the errors in the period determination are between 2 and 9 hours. The observed periodic variations of the brown dwarf candidates as well as of the T Tauri stars are interpreted as modulation of the flux at the rotation period by magnetically driven surface features, on the basis of a consistency with v sini values as well as (R-i) color variations typical for spots. Furthermore, the temperatures even for the brown dwarfs in the sample are relatively high (>2800K) because the objects are very young. Therefore, the atmospheric gas should be sufficiently ionized for the formation of spots on one hand and the temperatures are too high for significant dust condensation and hence variabilities due to clouds on the other hand.
astro-ph
astro-ph
Rotational periods of very young brown dwarfs and very low-mass stars in Cha I1 V. Joergens2, M. Fern´andez3, J. M. Carpenter4 and R. Neuhauser2,5 ABSTRACT We have studied the photometric variability of very young brown dwarfs and very low-mass stars (masses well below 0.2 M⊙) in the Cha I star forming region. We have determined photo- metric periods in the Gunn i and R band for the three M6.5 -- M7 type brown dwarf candidates Cha Hα 2, Cha Hα 3 and Cha Hα 6 of 2.2 to 3.4 days. These are the longest photometric peri- ods found for any brown dwarf so far. If interpreted as rotationally induced they correspond to moderately fast rotational velocities, which is fully consistent with their v sin i values and their relatively large radii. We have also determined periods for the two M5 -- M5.5 type very low-mass stars B 34 and CHXR 78C. In addition to the Gunn i and R band data, we have analysed JHKS monitoring data of the targets, which have been taken a few weeks earlier and confirm the periods found in the optical data. Upper limits for the errors in the period determination are between 2 and 9 hours. The observed periodic variations of the brown dwarf candidates as well as of the T Tauri stars are interpreted as modulation of the flux at the rotation period by magnetically driven surface features, on the basis of a consistency with v sin i values as well as (R-i) color vari- ations typical for spots. Furthermore, the temperatures even for the brown dwarfs in the sample are relatively high (> 2800 K) because the objects are very young. Therefore, the atmospheric gas should be sufficiently ionized for the formation of spots on one hand and the temperatures are too high for significant dust condensation and hence variabilities due to clouds on the other hand. A comparison with rotational properties of older brown dwarfs shows that most of the acceleration of brown dwarfs takes place within the first 30 Myr or less. If magnetic braking plays a role this suggests that the disk dissipation for brown dwarfs occurs between a few Myrs and 36 Myr. Subject headings: Stars: activity -- stars: fundamental parameters -- stars: B 34, CHXR 73, CHXR 78C) stars: late-type -- stars: low-mass, brown dwarfs -- stars: rotation individual (Cha Hα 1 to 12, 1Based on observations obtained at the European Southern Observatory at La Silla in program 65.L-0629. 2Max-Planck-Institut fur Extraterrestrische Physik, Giessenbachstrasse 1, D-85748 Garching, Germany. Email: [email protected] 3Instituto de Astrof´ısica de Andaluc´ıa (CSIC), Apdo. 3004, E-18080 Granada, Spain. Email: [email protected] 4Department of Astronomy, MS 105-24, California In- stitute of Technology, 1201 East California Boulevard, Pasadena, CA 91125. Email: [email protected] 5Astrophysikalisches Institut der Universitat von Jena, Schillergasschen 2-3, D-07745 Jena, Germany. Email: [email protected] 1. Introduction A photometric monitoring campaign of bona fide and candidate brown dwarfs in the Cha I star forming cloud has been carried out in two filters in order to study the time dependence of their bright- ness and color. It is known that magnetically driven surface features (spots) of stars modulate the brightness of the star as it rotates (e.g. Bou- vier et al. 1993 and references therein). The pre- sented photometric observations of brown dwarfs are aimed at the study of such spot-driven vari- abilities in the substellar regime in order to test if brown dwarfs have magnetic spots. Furthermore, since surface features modulate the emitted flux 1 at the rotational period, the photometric study is aimed at the determination of rotational periods for brown dwarfs. Rotational periods are fundamental (sub)stellar properties. The knowledge of rotational periods for objects covering a wide range of ages is im- portant for the understanding of the evolution of angular momentum. Scanning the period -- age -- diagram observationally for brown dwarfs with ages of less than 100 Myr provides in addition a test of substellar evolutionary theories since the onset of deuterium burning is expected to have an observable effect on the angular momentum evolu- tion. The contraction of brown dwarfs is expected to be temporarily decelerated or even stopped in the first several million years of their lifetime due to the ignition of deuterium (e.g. Burrows et al. 2001). However, since age estimates are also often model dependent, the significance of such a test might be somewhat limited. Furthermore, rota- tion rates are critical parameters for rotationally induced phenomena, like dynamo activity (sup- posed to cause surface spots) and meteorological processes. For solar-type main sequence stars there is a correlation between rotation and activity: the faster the rotation the more active the star is, measured in terms of chromospheric Hα and Ca II emission, flare activity as well as coronal X-ray emission. There are indications that this relation becomes invalid for later spectral types (late M) and lower masses. In particular near and below the substellar limit several rapid rotators are found with no or very little signs of chromospheric activ- ity (e.g. Basri & Marcy 1995; Delfosse et al. 1998; Gizis et al. 2000). For very young stars on the pre-main sequence (T Tauri stars) the relation between activity and rotation is still a matter of debate: correlations between rotation and X-ray emission have been found for T Tauri stars in Taurus (Bouvier 1990; Neuhauser et al. 1995; Stelzer & Neuhauser 2001), whereas a recent publication by Feigelson et al. (2002) reports the absence of a connection between X-ray emission and rotation for a large sample of T Tauri stars in Orion. Hα emission on the other hand is not a definite activity indicator for very young objects, because they often have circumstel- lar accretion disks, which are significant additional Hα emission sources. A study of the rotation- 2 activity-relation in the substellar regime at this very young age is hampered up to now by the lack of observational constraints of rotation parame- ters: v sin i values are known for two very young brown dwarfs and twelve brown dwarf candidates (Joergens & Guenther 2001, White & Basri 2003). Furthermore, Bailer-Jones & Mundt (2001) found photometric periods for two very young late-M dwarfs in σ Ori, which might be rotational peri- ods. For very cool objects (spectral type late M, L) surface spots might not play a significant role but another important process may affect the time de- pendence of the observed flux of the objects: below a temperature of about 2800 K the condensation of dust sets in (e.g. Tsuji et al. 1996a,b; Allard et al. 1997; Burrows & Sharp 1999). Inhomogeneities in dust cloud coverages may cause observable photo- metric variations. In spite of the demonstrated significance of rotational periods, the number of brown dwarfs with known rotational periods is rather small (see Sect. 2). This is particularly the case at very young ages. We have therefore carried out a photomet- ric monitoring campaign of brown dwarfs and very low-mass stars in the Cha I star forming cloud and determined periods for three brown dwarf candi- dates. The targets are introduced in Sect. 3, the data acquisition and analysis is described in Sect. 4 and 5. The results are presented in Sect. 6 and dis- cussed in Sect. 7, followed by a summary in Sect. 8. 2. Hitherto known periods for brown dwarfs Several studies of the photometric behavior of late-M and L-dwarfs have been carried out so far. They have led to the detections of a handful of brown dwarfs showing periodic variabilities and of many, mostly substellar L-dwarfs, showing non- periodic variabilities. We compiled all photomet- ric periods that we have found for brown dwarfs in the literature in Table 1 and ordered them by increasing age. The nature of the detected periodic variations are not finally clarified. Tinney & Tolley (1999) report variations of an M9 dwarf in narrow band filters, which are sensitive to changes in TiO ab- sorption features and therefore to clouds. Bailer- Jones & Mundt (2001) detected significant peri- ods below one day for four late-M and L dwarfs, among them the very young object S Ori 31. Fur- thermore, they have found hints for a period for S Ori 33. Out of these five periods four might be rotational periods according to the authors. The detected periods below one day for S Ori 31 and S Ori 33 imply very rapid rotational velocities of the order of 100 km s−1 (cf. Sect. 7.3) if confirmed as rotational periods. Furthermore, they find non- periodic variations for several L-dwarfs with time scales of hours and suggest that one sees the for- mation and dissipation of clouds rather than ro- tationally driven features. Mart´ın, Zapatero Os- orio & Lehto (2001) report a varying periodicity found in I band photometry of the M9.5 dwarf BRI 0021-0214. The authors suggest that the vari- ations are caused by inhomogeneous clouds rather than spots because the object is very inactive in terms of Hα and X-ray emission despite a very fast rotation. Clarke et al. (2002) found peri- odic variations of the brightness of Kelu-1. The detected period of 1.8 h is consistent with the ro- tational velocity. Nevertheless, Clarke & Tinney (2002) found no evidence for variability in dust sensitive molecular lines and the nature of the vari- ations remains still unclear. Recently, Gelino et al. (2002) found photometric variabilities for several L-dwarfs. These features show no periodicities or at least no persistent periodicities and are there- fore rather caused by rapid evolution of atmo- spheric features than being rotationally induced. Eisloffel & Scholz (2001) studied the photometric behavior of very low mass stars and brown dwarfs in the young (∼36 Myr) cluster IC4665 and report the finding of rotational periods below one day for five candidate brown dwarfs. Details on peri- ods and amplitudes will be given in a forthcoming publication (pers. comm.). 3. Sample The targets of our observations are twelve low- mass M6 -- M8 -- type objects, Cha Hα 1 to 12, lo- cated in the center of the Cha I star forming cloud with an age of 1 -- 5 Myr (Comer´on, Rieke & Neuhauser 1999; Comer´on, Neuhauser & Kaas 2000). Their masses are below or near the border line separating brown dwarfs and very low-mass stars. Four of them are bona fide brown dwarfs (Neuhauser & Comer´on 1998, 1999; Comer´on et al. 2000). We like to note, that all three brown dwarf candidates, for which we are presenting ro- tational periods in this paper, Cha Hα 2, 3 and 6, have masses below the hydrogen burning mass limit but they are candidates because the error bars of their effective temperatures extend into the stellar regime (cf. Comer´on et al. 2000). A first study of their rotational properties has been carried out by Joergens & Guenther (2001), who measured the rotational broadening of spec- tral features in high-resolution UVES spectra. They found v sin i values within the range of 8 to 26 km s−1 for nine out of the twelve objects. Furthermore, the very low-mass T Tauri stars B 34, CHXR73 and CHXR 78C, in the same field- of-view, have been studied. 4. Photometry We monitored a 13′ × 13′ region in the Cha I cloud photometrically in six consecutive half nights with DFOSC at the Danish 1.5 m tele- scope at ESO, La Silla, Chile. Images have been obtained in the Bessel R and the Gunn i filter between 2000 May 31 and June 5. The first two nights have been partly cloudy and therefore fewer images (with lower S/N) have been taken in these two nights. The exposure times have been chosen to detect periods on time scales of expected rotational peri- ods of the objects. Projected rotational velocities v sin i (Joergens & Guenther 2001) indicate that their rotational periods are within the range of a few days. The objects span a large dynamical range (I=13.6 to 17.4 mag) therefore we obtained Gunn i band images with two different exposure times of about 400 s and about 900 s. R band im- ages have been taken with only one exposure time of about 1000 s. We performed aperture photometry for the twelve bona fide and candidate brown dwarfs Cha Hα 1 to 12, the very low-mass T Tauri stars CHXR 73, CHXR 78C, B 34 as well as for several reference stars in the field with IRAF6. The sky background was determined from a source-free an- nulus and subtracted from the object counts. The 6IRAF is distributed by the National Optical Astronomy Observatories, which is operated by the Association of Uni- versities for Research in Astronomy, Inc. (AURA) under cooperative agreement with the National Science Founda- tion. 3 T Tauri stars CHXR 74 and Sz 23 are also in the field of view but have been saturated. The cal- culation of differential magnitudes allowed us to compensate for variable atmospheric conditions at least to a certain degree. Three reference stars have been chosen carefully for each filter based on the criteria of constant brightness over the time of observations, good S/N as well as similar bright- ness as the targets. From the analysis of these reference stars we estimate the photometric error to be about 0.015 mag or less: the standard devia- tion of the reference stars is 0.009 to 0.014 mag in the Gunn i band and 0.005 to 0.006 mag in the R band (cf. Fig. 1 and Fig. 3 to 6 for the dispersion of the reference stars). 5. Time series analysis We applied the string-length method (Dworet- sky 1983) in order to search for periodicity in the obtained light curves of the targets. This method is ideally suited for unevenly spaced data as in our case. The algorithm phase folds the data with a trial period and calculates the string length be- tween successive data points. This is done for all periods within a given period range. The period which generates the minimum string length is the most likely period. The significance of the detected periods was es- timated by cross checking a randomized data set sampled with the same time steps as the real data but with arbitrary magnitudes within the limits of the real magnitudes. For each suspected period, we checked 10000 randomized data sets. The per- centage of samples that have a longer (i.e. less significant) string length for any period than that of the suspected period yields the confidence level (a 99.99% confidence level corresponds to the case that all randomized samples have a longer string length than the string length for the suspected pe- riod). Besides this significance test, a major em- phasis was put on a direct check by eye of the original as well as the phase folded light curves. It is well known that in addition to intrinsic pe- riodicities of the monitored objects the light curves may show alias periods due to the sampling rate of the data. The most common alias periods Pfalse are P −1 false = 1.0027 d−1 ± P −1 true (1) period with possible alias periods, which are in- herent to the data due to an observing frequency of one sidereal day. In general, more images have been obtained in the Gunn i band than in the R band and in ad- dition the Gunn i band data have a higher S/N. Therefore for each object, firstly the Gunn i band data were studied and it was given them a higher weight. We have searched for periods in the range of 1.5 hours to 5 days. The minimum period is set by twice the average sampling rate of about 45 min. The maximum is chosen to be slightly smaller than the total time coverage of about 5.5 d. 6. Results We have measured significant periods for three brown dwarf candidates (Cha Hα 2, 3, 6) as well as for two very low-mass T Tauri stars (B 34, CHXR 78C) within the range of 2.2 d to 4.8 d and with Gunn i band amplitudes between 0.06 mag and 0.14 mag (Table 2 and Fig. 1, 3 -- 6). For Cha Hα 4, 5, 8, 12 and CHXR 73 no clear variations are detected and limits for variability amplitudes for these objects are derived (Table 2). The S/N of Cha Hα 1, 7, 9, 10 and 11 in the ob- tained images is too low for a further exploitation. These objects should be re-observed with a larger telescope. Details of the results of the period analysis on individual objects as well as original and phase folded light curves are presented in the follow- ing subsections. Objects are ordered by decreas- ing mass: from about 0.12 M⊙ for B 34 to about 0.05 M⊙ for Cha Hα 6. These masses are rough estimates based on a comparison with theoretical evolutionary tracks by Baraffe et al. (1998) for the Cha Hα objects (Comer´on et al. 2000) and by Bur- rows et al. (1997) for the T Tauri stars B 34 and CHXR 78C (Comer´on et al. 1999), respectively. Recently, a near-infrared (JHKS) photometric monitoring campaign of the Cha I cloud has been carried out (Carpenter et al. 2002), which in- cludes all the targets of our sample. These ob- servations contain between 11 and 25 points per target obtained in 10 -- 12 separate nights, mostly in 2000 April and May and for some targets (B 34, CHXR 78C and Cha Hα 3) there is an additional point in 2000 January. (Dworetsky 1983). This equation relates the true After checking that the period analysis per- 4 formed for our data clearly excludes periods shorter than one day, we have decided to take the data of Carpenter et al. (2002) into account for our analysis. Our Gunn i and R band light curves are well-sampled but have a total time ba- sis of less than six days and are therefore hardly providing strong support for periods longer than three days. On the other hand, the JHKS data include just one or two points per night but cover several weeks. Therefore, they are the ideal com- plement to our data set and increase the total time basis to 55 -- 59 days (not taking into account the single measurement from 2000 January). Since the two data sets correspond to different filters, we have joined only data taken with the Gunn i and J filters, for which the difference in ef- fective wavelength is smaller. In order to deal with the differences in amplitude and average bright- ness of the objects in the two different bands, we have normalized each data set by subtracting its mean and dividing by its standard deviation7. Thereafter, we have carried out a period analy- sis of the objects in the so-created joined Gunn i and J data set. The results are given for each object at the end of the corresponding subsection. Upper limits for the error of the period determi- nations have been computed from Nyquist's fre- quency based on a time basis of 55 to 59 days. They range from 9 to 2 hours. The error bars of the J band data, except for B 34, are between a fac- tor 0.3 and 0.4 of the amplitude observed in that band. For B 34 the error bars are smaller, namely between a factor 0.15 and 0.23 of its J amplitude. Only for the plots, the normalized iJ light curves have been additionally divided by their am- plitudes for clarity. 6.1. B 34 (∼ 0.12 M⊙) The brightness of the T Tauri star B 34 varies with a peak-to-peak amplitude of 0.14 mag in the Gunn i band and 0.18 mag in the R band during the six nights of our observations (Fig. 1). The light curves show a smooth variability apart from one runaway data point in the 3rd night, which is present in both filters and shows a larger de- 7If normalization is done by subtracting the middle value of the light curve and dividing by its amplitude, the same results are obtained but, mathematically, the first method is more consistent. viation in the R band than in the Gunn i band, which is consistent with an intrinsic increase of the brightness due to a flare-like event. The period search analysis of the Gunn i light curve of B 34 yields a clear period of 4.5 d with a confidence level above 99.99%. Searching for peri- ods in the R band data yields rather a plateau with a range of 4.2 d to 4.9 d for the string length than a clear minimum, i.e. the period search algorithm is unable to distinguish between the significance of periods within this range. This can be under- stood if one takes into account that these periods are very close to the total time basis of the data set (∼5.5 d) and that the R band brightness of the first two nights is only constrained by one data point each. Therefore, any errors of these points greatly affect the period search. The period analysis carried out on the Gunn i and J joined data set solves the uncertainty of the mentioned plateau and gives 4.75±0.38 d as the best period. The data folded with that period are shown in Fig. 1 (bottom). The phase folded (R-i) color curve (Fig. 2) shows that B 34 is redder during minimum light as expected for brightness variations caused by star spots. 6.2. CHXR 78C (∼ 0.09 M⊙) The T Tauri star CHXR 78C changes its bright- ness with an amplitude of 0.07 mag in Gunn i and 0.10 mag in R (Fig. 3). The period search algo- rithm returns a period of 4.0 d for the Gunn i band light curve and a period of 3.9 d for the R band light curve, both with a confidence level above 99.99%. The period analysis carried out on the Gunn i and J joined data set yields a slightly longer period of 4.25±0.31 d. The folded Gunn i and R band data confirm it as the best period (Fig. 3, bottom). 6.3. Cha Hα 2 (∼ 0.07 M⊙) In order to increase the S/N for the faint brown dwarf candidate Cha Hα 2 we binned adjacent data points in groups of about five. Fig. 4 (top and middle panel) displays the original data as well as the overplotted binned data. The binned light curves have variability amplitudes of 0.05 mag in Gunn i and 0.06 mag in R. We have found a min- imum string length for a period of 2.8 d for the Gunn i band data with a confidence level larger 5 than 99%. The light curves folded with this period support this result. The data analysis yields also the two less significant periods 0.73 d and 1.56 d, which cannot be rejected at once by a look to the phased light curves. However, they are related with the more significant 2.8 d period by Eq. 1 and are therefore supposed to be alias periods in- herent to the data due to the sampling frequency of one day. The R band light curve of Cha Hα 2 shows similar variations and trends as the Gunn i band light curve (Fig. 4, middle). However, the period analysis of the R band data does not con- firm the 2.8 d period on a high confidence level. It gives a minimum string length for 1.6 d (the alias in the Gunn i band) and only a second minimum for 2.8 d. The enhancement of the alias compared to the true period can be attributed to the smaller number of R band images as well as their lower S/N. The period analysis for this target carried out on the Gunn i and J joined data set gives 3.21±0.17 days as the best period. The Gunn i and R band data folded with this period show a smooth curve (Fig. 4, bottom). We note that there are hints in HST images of Cha Hα 2 that it may be a close 0.2′′ binary (Neuhauser et al. 2002) unresolved in our DFOSC aperture photometry. If this is confirmed, the pe- riod found for Cha Hα 2 is presumably the period of one of the two components or a combination of both. 6.4. Cha Hα 3 (∼ 0.06 M⊙) The brown dwarf candidate Cha Hα 3 exhibits variations with a Gunn i amplitude of about 0.08 mag and an R amplitude of about 0.09 mag (Fig. 5). The Gunn i band flux of this object seems to be modulated with a clear sine wave. However, the first three data points in the second night deviate from this behavior. The period analysis without these three points yields a highly signifi- cant period of 2.2 d with a confidence level above 99.99%. This suggests that 2.2 d is an intrin- sic period to this object, although strong reasons have not been found to reject the three deviant data points. The R band data reflect the general trends also seen in the Gunn i band but are more noisy. They support a period of 2.2 d but do not confirm it with high significance. The R band data of nights 3 to 6 show a minimum string length for 6 a period of 0.06 d but this can be rejected on the basis of a check of the data folded with this pe- riod. There is a second minimum at 2.2 d and the R band data folded with 2.2 d look quite smooth. The period analysis carried out on the Gunn i and J joined data set yields a period of 2.19 d ± 0.09 d for Cha Hα 3, after discarding one single outlying J data point. The light curves of this object folded with the derived period also confirm the result and are displayed in Fig. 5 (bottom). The outlying J data point was included again for the plot for completeness. It is the lowest one in the joined iJ light curve at a phase of about 0.55. 6.5. Cha Hα 6 (∼ 0.05 M⊙) In order to increase the S/N of the brown dwarf candidate Cha Hα 6, we binned adjacent data points in groups of two or three. See Fig. 6 (top and middle panel) for the original data and the overplotted binned data. The binned data of the object shows variabilities with an amplitude of 0.06 mag in both bands. The study of periodicities for the Gunn i band data results in a minimum string length for a period of 3.49 d with a confi- dence level of 96%. There are two less significant minima at 0.77 d and 1.43 d, which are related to the 3.49 d period by Eq. 1. The analysis of the R band data is hampered by the small number of data points. Particularly in the first night there is only a single data point which has a large un- certainty due to the poor weather conditions. We ignored it for the further analysis of the R band data and the search for periods. We find a mini- mum string length for a period of 3.34 d and two less significant minima for 0.78 d and 1.58 d. The alias periods for 3.34 d are 0.77 d and 1.42 d, there- fore 0.78 and 1.58 may be aliases. The period analysis carried out for the Gunn i and J joined data set reproduces exactly the period that was obtained for the Gunn i and R band data, 3.36±0.19 days. The confidence level is 99.99%. The phase folded light curves of Cha Hα 6 also confirms it as the best period (cf. Fig. 6 bot- tom). 7. Discussion The periods measured for Cha Hα 2, Cha Hα 3 and Cha Hα 6 are among the first photometric pe- riods determined for very young substellar objects and the longest found for any brown dwarf so far. Furthermore, the objects have relatively large variability amplitudes compared to hitherto mon- itored brown dwarfs. Causes of these variations as well as constraints for the evolution of angular momentum in the substellar regime are discussed in the following sections. 7.1. Rotational modulation due to spots We interpret the detected periodic photometric variations of the brown dwarf candidates as well as of the T Tauri stars as rotational modulation of the emitted flux due to surface spots based on the following arguments: (1) The detected periods are consistent with ro- tational velocities v sin i and are therefore likely rotational periods. See Sect. 7.2 for more details. (2) The monitored (R-i) colors of these objects show larger amplitudes for shorter wavelength in agreement with the expectations for spots. This is true for all but Cha Hα 6, which shows the same amplitude in both the Gunn i and R band. Nev- ertheless, there may be slight differences between the amplitudes in the two bands, which are swal- lowed by the measurements uncertainties. The J band amplitudes based on the data by Carpen- ter et al. (2002) for the same objects are al- ways smaller or similar than those observed in the Gunn i band also in agreement with spots. Only for Cha Hα 6, the J band amplitude is apparently slightly larger than the Gunn i amplitude. How- ever, the fact that the J band data have larger photometric errors than the R and Gunn i data could very well account for the larger J band am- plitude for Cha Hα 6. (3) Young, pre-main sequence stars are known to have dominant surface features (e.g. Bouvier et al. 1993 and references therein), which are at- tributed to magnetic dynamo action. Though the brown dwarfs as well as the T Tauri stars stud- ied in this paper have very low masses, they are relatively hot due to their young age. Thus, the atmospheric temperatures may be hot enough for a sufficient ionization fraction of the plasma in the atmospheres to allow for magnetically induced sur- face features. (It should be noted that the dynamo operating in fully convective objects, like pre-main sequence stars, stars of lower mass than ∼0.4 M⊙ and brown dwarfs, is not yet completely under- stood.) (4) The temperatures of these young objects are too hot (between 3030 K for B 34 and 2840 K for Cha Hα 3 and 6) for significant dust condensation according to model atmosphere calculations (e.g. Tsuji, Ohnaka & Aoki 1996a; Tsuji et al. 1996b; Allard et al. 1997; Burrows & Sharp 1999). There- fore, non-uniform condensate-coverage that has been suggested to explain the photometric vari- ability detected for some L and M type old and hence cool (brown) dwarfs (Tinney & Tolley 1999; Bailer-Jones & Mundt 2001; Mart´ın et al. 2001; Clarke et al. 2002) is an unlikely cause for the variabilities detected by us for the studied Cha I objects. 7.2. Comparison with spectroscopic veloc- ities v sin i As mentioned above, the detected periods for Cha Hα 2, 3, 6 and B 34 are consistent within the measurements uncertainties with their pro- jected rotational velocities v sin i of 12.8 km s−1, 21.0 km s−1, 13.0 km s−1 and 15.2 km s−1, respec- tively (Joergens & Guenther 2001). This indi- cates that they are likely rotational periods of the objects (for CHXR 78C no v sin i measurement is available). Rotational periods Pv sin i, which are upper limits of the true rotational periods have been derived from radii and v sin i values and are given in Table 2. The radii have been determined based on luminosities and effective temperatures from Comer´on et al. (1999, 2000) applying the Stefan-Boltzmann law. It is noteworthy that the observed periods Pphot are always -- with the exception of Cha Hα 2 - larger than the calculated periods Pv sin i but do still agree with them when the errors are taken into account, hinting at a systematic effect. A 1 σ error of Pv sin i with propagated errors of the luminosity, the effective temperature (∆ log Lbol/L⊙ = 0.175, ∆Teff = 150 K, cf. Comer´on et al. 2000) and v sin i (∆v sin i =1 -- 3 km s−1, cf. Joergens & Guen- ther 2001) is given in Table 2. Pv sin i scales with √Lbol, T −2 eff and v sin i−1. Reasons for systematic deviations of the estimated Pv sin i from the ob- served Pphot may be either that the luminosities (bolometric correction, distance) have been sys- tematically underestimated; the rotational veloci- ties v sin i have been systematically overestimated; the effective temperatures have been systemati- 7 cally overestimated and/or the radii are system- atically too small. 7.3. Radii of brown dwarfs The radii of the very young brown dwarfs and brown dwarf candidates Cha Hα 1 to 12 estimated from luminosities and effective temperatures are ranging from 0.3 R⊙ for Cha Hα 11 (∼0.03 M⊙) to 0.9 R⊙ for Cha Hα 4 (∼0.1 M⊙). The uncertainties are of the order of 30% due to propagated errors in the luminosities and effective temperatures. Their relatively large radii are fully consistent with the extremely young age of the objects and the fact that they are still in a contracting stage. Models of Chabrier & Baraffe (1997) show that a brown dwarf with a mass of 0.06 M⊙ has a radius of ∼0.55 R⊙ at 3 Myr and a radius of ∼0.45 R⊙ at 5 Myr in their calculations. Only after ∼500Myr, brown dwarfs have more or less shrunken to their final size at a radius of ∼ 0.1 R⊙ independent of their mass (Chabrier et al. 2000). Bailer-Jones & Mundt (2001) studied the pho- tometric variability of late-M and L dwarfs, among them several very young objects in σ Orionis. The authors claim that none of the objects in their sample has a radius larger than 0.2 R⊙. Based on this radius assumption and on spectroscopic veloc- ities v sin i of 10-60 km s−1, as found by Basri et al. (2000) for late-M and L dwarfs in the field, they in- fer expected rotational periods for their targets of the order of 1 to 10 hours (see also Sect. 7.4). How- ever, the radii for the σ Orionis objects in their sample, which have an age of 1 -- 5 Myr and masses between 0.02 and 0.12 M⊙, are certainly larger than 0.2 R⊙. This is not only shown by a com- parison with theoretical models (e.g. Chabrier et al. 2000) but there is also observational evidence for it from the study of the L1.5 dwarf S Ori 47. Estimations of its effective temperature and lumi- nosity (Zapatero Osorio et al. 1999) indicate that the radius even for this very low-mass brown dwarf (∼0.015 M⊙) is of the order of 0.35 R⊙. The S Ori objects studied by Bailer-Jones & Mundt (2001) are of significantly higher mass than S Ori 47 and have therefore most certainly much larger radii than it. 7.4. Evolution of angular momentum In the current picture of angular momentum evolution, young solar-mass stars as well as brown dwarfs are supposed to undergo a spin up due to contraction on the Hayashi track. Interaction with an accretion disk can hold up the acceleration of the rotation by magnetic braking until the inner disk is dissipated; then the star begins to spin up. As the star ages on the main sequence, an- gular momentum loss through stellar winds spins down the star again. In contrast, at the very low masses of brown dwarfs, there is supposed to be no braking due to winds, which would explain the observed (very) fast rotation of old brown dwarfs, which have v sin i values ranging from 5 km s−1 up to 60 km s−1 (Basri & Marcy 1995; Mart´ın et al 1997; Tinney & Reid 1998; Basri et al. 2000). Fur- thermore, brown dwarfs with ages between about 36 Myr and 1 Gyr seem to have rotational periods shorter than one day (see Sect. 2 for references). The lower age limit at about 36 Myr is set by ro- tational periods determined for brown dwarf can- didates in the young cluster IC 4665 (Eisloffel & Scholz 2001). However, the periods that we have measured for three brown dwarf candidates in Cha I range be- tween 2 and 3 days and are therefore larger than any period reported up to now for such low-mass objects. The relatively long rotational periods and moderately fast rotational velocities of the brown dwarf candidates in Cha I are explained naturally within the current picture of angular momentum evolution by the fact that they are still in an early contracting stage. They may have furthermore suffered until very recently a braking due to their interaction with an accretion disk. The comparison of our rotation periods at 1 -- 5 Myr with those in the literature at > 36 Myr, gives first indications that most of the acceleration of brown dwarfs takes place in the first 30 Myr or less of their lifetime. It is known that Cha Hα 2 and 6 have optically thick disks (Comer´on et al. 2000), therefore magnetic braking due to interac- tions with the disk may play a role for them. This is suggested by the fact, that among the three brown dwarf candidates with determined periods, the one without a detected disk, Cha Hα 3, has the shortest period. If the interaction with the disk is responsible for the braking, our results and the one 8 from Eisloffel & Scholz (2001) indicate that brown dwarf inner disks dissipate between 1 -- 5 Myr and 36 Myr. These limits for the time scale of disk dis- sipation for brown dwarfs are not inconsistent with that for T Tauri stars, which are known to dissolve their inner disks within about the first 10 Myr (e.g. Calvet et al. 2000). Bailer-Jones & Mundt (2001) report the detec- tion of photometric periods for the two very young (1 -- 5 Myr) M6.5 dwarfs S Ori 31 and S Ori 33 of less than 9 h (cf. Table 1) and suggest that these might be rotational periods of the objects. However, as described in Sect. 7.3, a radius of <0.2 R⊙ is un- reliable for the young S Ori objects; a value of the order of half to one solar radius is more consis- tent (e.g. Chabrier et al. 2000). If the periods of less than 9 h for S Ori 31 and S Ori 33 are con- firmed as rotational periods, their rotational veloc- ities would be of the order of 100 km s−1, assum- ing for example a radius of 0.6 R⊙. This would mean that these very low-mass stars rotate at a much higher speed than the brown dwarfs and brown dwarf candidates in Cha I, which have spec- troscopic velocities v sin i of 9 -- 28 km s−1 (Joergens & Guenther 2001) and rotational periods of 2 to 3 days as shown in this paper. A possible explana- tion for a much faster rotation of the S Ori objects compared to the brown dwarfs in Cha I despite the similar age, would be a significant difference in their disk life-times, at least if magnetic brak- ing due to a circumstellar disk is an important process for them. The dissipation of circumstel- lar material in σ Orionis could be enhanced due to strong winds from the hot OB stars in this re- gion in comparison with the Cha I region without such a hot star. Therefore disk life-times could be much shorter in σ Orionis leading to an early spin up and, thus, to significantly faster rotation of the S Ori objects than of the brown dwarfs and very low-mass stars in Cha I. A measurement of v sin i of S Ori 31 and S Ori 33 would be useful to confirm their proposed very fast rotation. However, high-resolution spectroscopy for these faint objects is challenging even with large telescopes. 8. Summary We have determined photometric periods in the Gunn i and R band for the three M6.5 -- M7 type brown dwarf candidates Cha Hα 2, Cha Hα 3 and Cha Hα 6 of 3.2 d, 2.2 d and 3.4 d, respec- tively. These are the longest photometric peri- ods found for any brown dwarf so far. If in- terpreted as rotationally induced they correspond to moderately fast rotational velocities, which is fully consistent with their v sin i values (Joergens & Guenther 2001) and their relatively large radii. Furthermore, we have detected periods for the two M5 -- M5.5 type very low-mass stars B 34 and CHXR 78C of 4.8 d and 4.3 d. For the determi- nation of the periods, we took in addition to our optical Gunn i and R photometry also J band data obtained by Carpenter et al. (2002) into account. Our Gunn i and R band data are sampled with a relatively high frequency (∼0.5 h) but have a relatively short time base (5.5 d), whereas the J band data consist of 10 -- 24 data points spread over 41 nights immediately before our observations and are therefore ideally complementing our data and increasing the total time coverage to 55 -- 59 days. The observed periodic variations of the brown dwarf candidates as well as of the T Tauri stars are interpreted as modulation of the flux at the rotation period due to magnetically driven surface features on the basis of a consistency with v sin i values as well as (R-i) color variations typical for spots. Furthermore, the temperatures even for the brown dwarfs in the sample are relatively high (> 2800 K) because the objects are very young. Therefore, the atmospheric gas should be suffi- ciently ionized for the formation of spots on one hand and the temperatures are too high for signifi- cant dust condensation and hence variabilities due to clouds on the other hand. These first indica- tions for surface spots on very young brown dwarfs support the overall picture of magnetic activity of brown dwarfs, which is emerging in the last years: old (cool) brown dwarfs tend towards an absence of persistent activity, whereas very young brown dwarfs seem to have more in common with T Tauri stars as they are relatively active in terms of X- ray emission (e.g. Neuhauser et al. 1999) and Hα emission. Estimation of the radii for the bona fide and candidate brown dwarfs Cha Hα 1 to 12 show that they are relatively large (0.3 -- 0.9 R⊙) for such low- mass objects. They are fully consistent with the- oretical calculations of the contraction time scales of such very young brown dwarfs and objects near 9 the hydrogen burning mass limit. The periods measured by us provide valuable data points in an as yet, in terms of rotational characteristic, almost unexplored region of the age-mass diagram. A comparison of the deter- mined rotational periods at the age of a few mil- lion years with rotational properties of older brown dwarfs (>36 Myr, Eisloffel & Scholz 2001) shows that most of the acceleration of brown dwarfs dur- ing the contraction phase takes place within the first 30 Myr or less and suggests that the disk dis- sipation for brown dwarfs occurs between 1 -- 5 Myr and 36 Myr. However, a larger sample of brown dwarfs at various ages and in particular at a few million years with known rotation rates is much needed to further constrain the substellar angular momen- tum evolution as well as to allow for statistics of rotation rates in connection with activity indica- tors. We like to acknowledge helpful discussions on the topic of this paper with G. Wuchterl, B. Stelzer and C. Broeg. Furthermore, we thank our referee, C. Bailer-Jones, for valuable com- ments, which helped to improve the paper. We are also grateful to E. Guenther for contributions to the photometric monitoring campaign, R. Gar- rido for his help concerning the uncertainties of the period determination and to the ESO staff at La Silla for their support during the DFOSC observations. VJ acknowledges grant from the Deutsche Forschungsgemeinschaft (Schwerpunkt- programm 'Physics of star formation'). MF was partially supported by the spanish grant PB97- 1438-C02-02. RN acknowledges financial support from the BMBF through DLR grant 50OR 0003. REFERENCES Allard, F., Hauschildt, P. H., Alexander, D. R. & Starrfield S. 1997, ARA&A, 35, 137 Bailer-Jones, C. A. L. & Mundt, R. 2001, A&A, 367, 218 Baraffe, I., Chabrier, G., Allard, F. & Hauschildt, P. H. 1998, A&A, 337, 403 Basri, G. & Marcy, G. W. 1995, AJ, 109, 762 Basri, G., Mohanty, S., Allard, F., Hauschildt, P. H., Delfosse, X., Mart´ın, E. L., Forveille, T. & Goldman, B. 2000, ApJ, 538, 363 Bouvier, J. 1990, AJ, 99, 946 Bouvier, J., Cabrit, S., Fern´andez, M., Mart´ın, E. L. & Matthews, J. M. 1993, A&A, 272, 176 Burrows, A. et al. 1997, ApJ, 491, 856 Burrows, A. & Sharp, C. M. 1999, ApJ, 512, 843 Burrows, A., Hubbard, W. B., Lunine, J. I. & Liebert J. 2001, Rev. Mod. Phys. 73, 3 Carpenter, J.M., Hillenbrand, L.A., Skutskie, M.F., & Meyer, M.R. 2002, AJ 124, 1001 Chabrier, G. & Baraffe, I. 1997, A&A, 327, 1039 Chabrier, G., Baraffe, I., Allard, F. & Hauschildt, P. 2000, ApJ, 542, 464 Clarke, F. J. & Tinney, C. G. 2002, Poster pre- sented at the IAU Symp. 211, Brown Dwarfs, Hawai'i, May 20-24, 2002 Clarke, F. J., Tinney, C. G. & Covey, K. R. 2002, MNRAS, 332, 361 Comer´on, F., Rieke, G. H. & Neuhauser, R. 1999, A&A, 343, 477 Comer´on, F., Neuhauser, R. & Kaas, A. A. 2000, A&A, 359, 269 Delfosse, X., Forveille, T., Perrier, C. & Mayor, M. 1998 A&A, 331, 581 Dworetsky, M. M. 1983, MNRAS, 203, 917 Eisloffel, J. & Scholz, A. 2001, in ESO Astro- physics Symp., The Origins of Stars and Plan- ets: The VLT View, Garching, Germany, April 24-27, 2001, ed. J. F. Alves & M. J. McCaugh- rean, 127 Feigelson, E. D., Broos, P., Gaffney, J. A. III, Garmire, G., Hillenbrand, L. A., Pravdo, S. H., Townsley, L. & Tsuboi, Y. 2002, ApJ, 574, 258 Gelino C.R., Marley M.S., Holtzman J.A., Acker- man A.S. & Lodders K. 2002, ApJ 577, 433 Gizis, J. E., Monet, D. G., Reid, I.N., Kirkpatrick, J. D., Liebert, J. & Williams, R. J. 2000, AJ, 120, 1085 10 Joergens, V., Guenther, E. 2001, A&A, 379, L9 Mart´ın, E. L., Basri, G., Delfosse, X. & Forveille, T. 1997, A&A, 327, L29 Mart´ın, E. L., Zapatero Osorio M. R. & Lehto H.J. 2001, ApJ, 557, 822 Neuhauser, R., Sterzik, M. F., Schmitt, J. H. M. M., Wichmann, R. & Krautter, J. 1995, A&A, 297, 391 Neuhauser, R. & Comer´on, F. 1998, Science, 282, 83 Neuhauser, R. & Comer´on, F. 1999, A&A, 350, 612 Neuhauser, R. et al. 1999, A&A, 343, 883 Neuhauser, R., Brandner, W., Alves, J., Joergens, V. & Comer´on, F. 2002, A&A, 384, 999 Stelzer, B. & Neuhauser, R. 2001, A&A, 377, 538 Terndrup, D. M., Krishnamurthi, A., Pinson- neault, M. H. & Stauffer, J. R. 1999, AJ, 118, 1814 Tinney, C. G. & Reid, I. N. 1998, MNRAS, 301, 1031 Tinney, C. G. & Tolley, A. J. 1999, MNRAS, 304, 119 Tsuji, T., Ohnaka, K. & Aoki W. 1996a, A&A, 305, L1 Tsuji, T., Ohnaka, K., Aoki, W. & Nakajima T. 1996b, A&A, 308, L29 White, R. J. & Basri, G. 2003, ApJ, 582, 1109 Zapatero Osorio, M. R., B´ejar, V. J. S., Rebolo, R., Mart´ın, E. L. & Basri, G. 1999, ApJ, 524, L115 This 2-column preprint was prepared with the AAS LATEX macros v5.0. 11 6 7 8 9 10 11 MJD - 51690 6 7 8 9 10 11 MJD - 51690 0.6 0.7 0 0.5 1 1.5 2 Fig. 2. -- Phase folded (R-i) colors of B 34. -1.1 -1 -0.3 -0.4 -0.4 -0.3 -0.4 -0.3 -1 0 1 -0.4 -0.3 0 0.5 1 Phase 1.5 2 Fig. 1. -- Light curves of the very low-mass T Tauri star B 34. Top and middle panel: Gunn i and R band original light curves (relative magni- tudes), each with reference star magnitudes plot- ted below for comparison. Bottom panel: joined iJ and R light curves phase folded with the deter- mined period. Filled circles: our Gunn i and R 12 -0.7 -0.6 -0.3 -0.4 0.3 0.4 -0.4 -0.3 -1 0 1 0.3 0.4 6 7 8 9 10 11 MJD - 51690 6 7 8 9 10 11 MJD - 51690 0 0.5 1 Phase 1.5 2 3. -- Light curves of the very low-mass Fig. T Tauri star CHXR 78C. Panels are as in Fig. 1. 13 6 7 8 9 10 11 MJD - 51690 6 7 8 9 10 11 MJD - 51690 -0.3 -0.2 -0.3 -0.4 0.8 0.9 -0.4 -0.3 -1 0 1 0.8 0.9 0 0.5 1 Phase 1.5 2 Fig. 4. -- Light curves of the brown dwarf candi- date Cha Hα 2. Panels are as in Fig. 1. Filled cir- cles in the Gunn i and R light curves (upper and middle panel) represent an average of the original data points (open circles). 14 6 7 8 9 10 11 MJD - 51690 6 7 8 9 10 11 MJD - 51690 -0.5 -0.4 -0.3 -0.4 0.6 0.7 -0.4 -0.3 -1 0 1 0.6 0.7 0 0.5 1 Phase 1.5 2 Fig. 5. -- Light curves of the brown dwarf candi- date Cha Hα 3. Panels are as in Fig. 1. 15 -0.5 -0.4 -0.3 -0.4 0.5 0.6 -0.4 -0.3 -1 0 1 0.5 0.6 6 7 8 9 10 11 MJD - 51690 6 7 8 9 10 11 MJD - 51690 0 0.5 1 Phase 1.5 2 Fig. 6. -- Light curves of the brown dwarf candi- date Cha Hα 6. Panels are as in Fig. 1. Filled cir- cles in the Gunn i and R light curves (upper and middle panel) represent an average of the original data points (open circles). 16 Table 1: Photometric periods of brown dwarfs from the literature. Object Age Mass [M⊙] SpT Period S Ori 31 S Ori 33 CFHT-PL 8 LP 944-20 BRI 0021-0214 2M1334 Kelu-1 2M1146 SDSS 0539 2MASS 0746+20AB 2MASS 1300+19 1-5 Myr 1-5 Myr 100-120 Myr 475-650 Myr (M6.5) (M6.5) 0.08 . . . 0.06 M9 ≥1 Gyr ≥1 Gyr ≥1 Gyr ≥1 Gyr ≥1 Gyr ≥1 Gyr ≥1 Gyr M9.5 L1.5 <0.07 L2 L3 L5 L0.5 L1 7.5 h 8.6 h, 6.5 h 9.6 h . . . 19.2 h, 4.8 h 2.7 h, 6.3 h, 1.0 h 1.8 h 5.1 h 13.3 h 31.0 h 238. h Amplitude [mag] 0.012 0.010 0.028 0.04∗ 0.018, 0.007 0.020 0.012∗ 0.015 0.011 0.010 0.015 ref 1 1 2 3 4 1 5 1 1 6 6 Note. -- S Ori 31, S Ori 33 and CFHT-PL 8 are eventually stars but have been included because of their closeness to the hydrogen burning mass limit. Masses are estimates. Ages of the field brown dwarfs BRI 0021-0214 to 2MASS1300 are unknown but supposed to be of the order of 1 Gyr or more. The spectral types of S Ori 31 and S Ori 33 are estimates. Amplitudes are mea- sured in broad band I filters, except for the ones marked with an asterisk, these refer to narrow band observations (cf. references for details). Some object names are acronyms as given by the authors; their full names are: S Ori 31 = S Ori J053820.8-024613; S Ori 33 = S Ori J053657.9-023522; CFHT-PL 8 = VLC J0342268+245021; 2M1334 = 2MASSW J1334062+194034; 2M1146 = 2MASSW J1146345+223053; SDSS 0539 = SDSSp J053951.99-005902.0; 2MASS 0746+20AB = 2MASSI J0746425+200032; 2MASS 1300+19 = 2MASSW J1300425+191235. References. -- (1) Bailer-Jones & Mundt 2001; (2) Terndrup et al. 1999; (3) Tinney & Tolley 1999; (4) Mart´ın et al. 2001; (5) Clarke et al. 2002; (6) Gelino et al. 2002. 17 Table 2: Rotational periods and photometric amplitudes for brown dwarfs and very low-mass stars in Cha I I Object Characteristics of the targets SpT M∗ [M⊙] 0.12 0.09 0.07 0.06 0.05 0.15 0.1 0.1 0.07 0.05 [mag] 14.3 14.8 15.1 14.9 15.1 15.6 14.3 14.7 15.5 15.6 R∗ [R⊙] 0.93 0.94 0.73 0.77 0.68 1.46 0.89 0.83 0.59 0.66 M5 B 34 CHXR 78C M5.5 M6.5 Cha Hα 2 M7 Cha Hα 3 Cha Hα 6 M7 M4.5 CHXR 73 M6 Cha Hα 4 M6 Cha Hα 5 Cha Hα 8 M6.5 M7 Cha Hα 12 Periods Phot. amplitude ∆ i Pphot ∆Pphot Pv sin i ∆Pv sin i ∆ R [mag] [mag] 0.14 0.18 0.07 0.10 0.05 0.06 0.08 0.09 0.06 0.06 <0.03 <0.05 <0.03 <0.03 <0.04 <0.02 <0.02 <0.04 <0.04 <0.03 [d] 0.38 0.31 0.17 0.09 0.19 . . . . . . . . . . . . . . . [d] 4.75 4.25 3.21 2.19 3.36 . . . . . . . . . . . . . . . [d] +1.7 −1.0 . . . +1.4 −1.0 +0.8 −0.6 +2.0 −1.0 . . . . . . . . . . . . . . . [d] 3.1 . . . 2.9 1.9 2.6 . . . . . . . . . . . . . . . ∆ J [mag] 0.13 0.06 0.05 0.08 0.10 . . . . . . . . . . . . . . . Note. -- Spectral types (SpT), masses (M∗) and I band magnitudes are from Comer´on et al. (1999, 2000). Masses are estimates based on Baraffe et al. (1998) tracks for the Cha Hα objects and on Burrows et al. (1997) tracks for the T Tauri stars, respectively. Radii (R∗) are estimated based on luminosities and effective temperatures from Comer´on et al. (1999, 2000); their errors are of the order of 30%. Rotational periods (Pphot) and photometric peak-to-peak amplitudes (∆ R, ∆ i, ∆ J) are derived from our Gunn i and R band light curves taking into account J band data from Carpenter et al. (2002). ∆Pphot is an upper limit for the error of the derived periods based on Nyquist's frequency. Pv sin i denotes rotational periods derived from radii and v sin i measurements (Joergens & Guenther 2001). ∆Pv sin i is a 1 σ error. Note that ∆ i refers to our Gunn i photometry, whereas the I band magnitudes by Comer´on et al. are obtained in the Cousins I filter. For CHXR 78C no v sin i measurements are available. See the text for more details. 18
astro-ph/0701205
3
0701
2007-08-11T02:15:32
On the Mechanism of Gamma-Ray Burst Afterglows
[ "astro-ph" ]
The standard model of afterglow production by the forward shock wave is not supported by recent observations. We propose a model in which the forward shock is invisible and afterglow is emitted by a long-lived reverse shock in the burst ejecta. It explains observed optical and X-ray light curves, including the plateau at 10^3-10^4 s with a peculiar chromatic break, and the second break that was previously associated with a beaming angle of the explosion. The plateau forms following a temporary drop of the reverse-shock pressure much below the forward-shock pressure. A simplest formalism that can describe such blast waves is the ``mechanical'' model (Beloborodov, Uhm 2006); we use it in our calculations.
astro-ph
astro-ph
On the Mechanism of Gamma-Ray Burst Afterglows Z. Lucas Uhm, Andrei M. Beloborodov1 Physics Department and Columbia Astrophysics Laboratory, Columbia University, 538 West 120th Street, New York, NY 10027. ABSTRACT The standard model of afterglow production by the forward shock wave is not supported by recent observations. We propose a model in which the forward shock is invisible and afterglow is emitted by a long-lived reverse shock in the burst ejecta. It explains observed optical and X-ray light curves, including the plateau at 103 − 104 s with a peculiar chromatic break, and the second break that was previously associated with a beaming angle of the explosion. The plateau forms following a temporary drop of the reverse-shock pressure much below the forward- shock pressure. A simplest formalism that can describe such blast waves is the "mechanical" model (Beloborodov & Uhm 2006); we use it in our calculations. Subject headings: gamma rays: bursts -- hydrodynamics -- radiation mecha- nisms:nonthermal -- relativity -- shock waves 1. INTRODUCTION Since its discovery, gamma-ray burst (GRB) afterglow has been attributed to the forward shock wave (M´esz´aros & Rees 1997). This hypothesis was consistent with first follow-up observations at tobs ∼ 0.1 − 10 days after prompt GRBs. However, recent observations of early afterglows (tobs < 0.1 day) are difficult to reconcile with the model (e.g. Zhang 2007). One outstanding problem is the low-level plateau in the X-ray light curve (Nousek et al. 2006); this and other problems are summarized in § 3, where we discuss their possible resolution. Besides the forward shock (FS), a reverse shock (RS) is expected in the burst ejecta. A short-lived RS was proposed to emit a brief optical flash (M´esz´aros & Rees 1999; Sari & Piran 1999), and later absence of such flashes was interpreted as a lack of RS emission 1Also at Astro-Space Center, Lebedev Physical Institute, Profsojuznaja 84/32, Moscow 117810, Russia -- 2 -- (Roming et al. 2006). A long-lived RS in stratified ejecta with a decreasing Lorentz factor Γej was proposed as a mechanism of gradual energy injection into the FS (Rees & M´esz´aros 1998) and the RS emission was invoked to explain radio data (Panaitescu & Kumar 2004). The interaction of power-law ejecta with a power-law medium obeys scaling laws that are derived analytically (Rees & M´esz´aros 1998). However, the general problem of explosions driven by ejecta with arbitrary stratification of Γej remained unsolved. We have developed a simple "mechanical" formalism for such explosions (Beloborodov & Uhm 2006; hereafter BU06), which allows us to explore a new class of dynamical models, with rapid and strong evolution of the RS. We calculate the emissions from FS and RS and search for a scenario that would reproduce the observed X-ray and optical light curves. We find that observations can be explained if it is only the RS that contributes to the observed afterglow and the FS is invisible because of its extremely low radiative efficiency. 2. AFTERGLOW MODEL The central explosion ejects a cold (adiabatically cooled) relativistic flow. It may be viewed as a sequence of shells of energy δEej that coast with Lorentz factors Γej. Each shell can be prescribed an ejection time τ at a small radius r, and functions dEej/dτ and Γej(τ ) completely describe the ejected flow. The ejecta density is derived from continuity equation, ρej(τ, r) = dEej/dτ 4πr2Γ2 ejc3 (cid:0)1 − rΓ′ ej/cΓ3 ej(cid:1) , (1) where Γ′ ej ≡ dΓej/dτ ≤ 0. The ejecta push the blast -- a thin shell of compressed gas between FS and RS -- as described by equations (11)-(16) in BU06. We solve these equations numerically, tracking self-consistently the location of RS in the ejecta. The blast-wave model is determined by three functions: ρ1(r) (external density), Γej(τ ), and Eej(τ ). When the condition dΓej/dτ ≫ cΓ3 ej/r is satisfied, equation (1) simplifies, ρej = Γej 4πr3c2 dΓej dEej (cid:12) (cid:12) (cid:12)(cid:12) −1 (cid:12) (cid:12) (cid:12)(cid:12) , r cΓ2 ej ≫ τ d log Γej d log τ (cid:12) (cid:12) (cid:12)(cid:12) −1 (cid:12) (cid:12) (cid:12)(cid:12) . (2) Then the ejection time τ drops out, and only one function Γej(Eej) describes the ejecta. Once the dynamical equations are solved, we calculate the synchrotron emission from the shocked medium (FS emission) and shocked ejecta (RS emission). This calculation uses the standard model with three parameters (e.g. Piran 2004): fraction ǫe of the shock energy that goes to electron acceleration, slope p of the electron spectrum, and magnetic parameter ǫB = (B2/8πU) < 1, where U is the energy density of the shocked gas. The blast is treated -- 3 -- as one body in the sense that it moves as one body, with a common Γ. However, we take into account that it is made of different hot shells that pile up from the FS and RS. The synchrotron emissivity of each shell is tracked as the blast expands. The sum of observed emissions from all shells is found taking into account the velocity and curvature of the shells. The main points of this work are demonstrated by the numerical model presented in Figures 1-2. It assumes a uniform external medium with n1(r) = ρ1/mp = 1 cm−3 and a total isotropic energy of the explosion Eb = 1054 ergs. The stratification function Γej(Eej) (Fig. 1) assumes a "head" of the ejecta with Γej = 300, which carries an energy ∼ Eb/3. The remaining 2/3 of energy is carried by a slower "tail." The RS crosses the head at early times tobs <∼ 102 s and proceeds to the tail. Equation (2) becomes valid in the tail. The emission produced by the FS with usual parameters ǫe = 0.1, ǫB = 0.01, and p = 2.3 is shown by the thin curves in Figure 2e. Both X-ray and optical light curves are inconsistent with observations and illustrate the problems of the FS model. The theoretical optical light curve peaks at 103 − 104 s, when the peak frequency of synchrotron emission passes through the optical band (e.g. Sari et al. 1998). This peak is not observed. The theoretical X-ray light curve has a long monotonic decay with slope α ∼ 1.1 By contrast, observed X-ray afterglows are much weaker at 102 − 104 s; they show an initial steep decay to a low emission level and then a plateau. The RS emission for the same explosion is shown by thick curves in Figure 2e. The same parameters ǫe = 0.1, ǫB = 0.01, and p = 2.3 are assumed for the RS. Then the cooling frequency νc is between optical and X-ray bands throughout most of the afterglow. The X-ray emission is produced by fast-cooling electrons near the RS. Its initial peak (marked A in Fig. 2e) is followed by the steep decay AB and the plateau BC. This behavior can be understood from Figures 2a-2d, which show the explosion dynamics. The ejecta density drops dramatically behind the head of the ejecta because of the large gradient of Γej (see Fig. 1 and eq. 2). Therefore, X-ray emission is strongly reduced when the RS enters the tail. Its decay at tobs = 102 − 103 s has a temporal index α ∼ 3 and is limited only by the spherical curvature of the RS. The X-ray emission does not recover until the RS propagates to the region of flatter stratification function (smaller dΓej/dEej and higher ρej). This recovery begins at tobs ∼ 103 s and corresponds to the beginning of the X-ray plateau (point B). During the plateau stage, ρej at the RS grows and reaches a maximum at point C where dΓej/dEej is near its minimum. This point is the end of the X-ray plateau. 1It deviates from the power-law because of energy injection into the FS from the tail. The growth of blast-wave energy (by a factor of 3) implies a deviation from the standard deceleration law Γ2ρ1r3 = const. -- 4 -- Fig. 1. -- Stratification function Γej(Eej) of the numerical model. Here Eej(τ ) is kinetic energy of the ejected flow. It serves as a Lagrangian coordinate: Eej = 0 corresponds to the first ejected shell and Eej = Eb corresponds to the last shell; Eb = 1054 ergs is the total energy of explosion. The reverse shock (RS) starts at Eej = 0 and moves toward Eej = Eb, passing through points A, B, C, and D. The transition from "head" to "tail" of the ejecta occurs between points A and B. Point C is close to the inflection point of function Γej(Eej) and corresponds to the maximum of ρej (Fig. 2). Point D is near the end of the ejecta. In the exact numerical model, the ejected flow is described by two functions, Γej(τ ) and Eej(τ ) with 0 < τ < τb ≈ 100 s. However, for the afterglow emitted after the RS passes point A, only one function is important -- Γej(Eej) -- and τ ≈ 0 can be assumed for all shells after point A (see the text and eq. 2). Therefore, we show only Γej(Eej) here. -- 5 -- Fig. 2. -- Panels (a)-(d) show the blast-wave evolution in the mechanical model. (a) Density of the preshock ejecta nej(RS) = ρej/mp as a function of radius r of the expanding blast wave; the assumed external density n1 = ρ1/mp = 1 cm−3 is shown for comparison. (b) Lorentz factors of the preshock ejecta, Γej(RS), and the blast, Γ. (c) Pressures at the FS and RS, pf and pr. (d) Relative Lorentz factor of the RS, γ43 (see BU06). Panels (e) and (f) show observed emission from the blast wave as a function of observer time. The emission parameters are ǫB = 0.01, ǫe = 0.1, and p = 2.3. (e) Observed spectral flux, assuming the burst is at a cosmological redshift z = 1. Thin curves show the FS emission -- X-ray (dashed line) and optical (dotted line). Thick curves show the RS emission -- X-ray (solid line) and optical (dash-dotted line). (f) Evolution of spectral indices βX (X-ray) and βO (optical) for the RS emission. The transition from the slow cooling spectrum β = (p − 1)/2 = 0.65 (νc > ν) to the fast cooling spectrum β = p/2 = 1.15 (νc < ν) occurs early in the X-ray band and later in the optical band. The peaks in βX and βO will be discussed elsewhere (Z. L. Uhm & A. M. Beloborodov 2007, in preparation). The correspondence between the marked points in dynamical panels (a)-(d) and emission panels (e)-(f) is approximate: radiation received -- 6 -- Following point C, the X-ray light curve has a slope α ∼ 1 until the last break at point D. This break corresponds to the RS reaching the end of the ejecta, and its observed time is tobs ∼ r/2Γ2 endc. In our example model, Γend ≈ 20 (Fig. 1), which gives the break D at tobs ∼ 105 s. (We also ran models with smaller Γend; then the power-law with α ∼ 1 extends to longer times ∼ weeks.) The steep slope of the light curve after break D is limited by the spherical curvature of the RS. The RS optical light curve also differs significantly from the prediction of the FS model. It has an initial power-law decay at 102 − 3 × 103 s followed by the hump or "shoulder." The optical light curve can be understood by noting that it is produced by slow-cooling electrons. When crossing the head of the ejecta, the RS creates a population of nonthermal electrons that continue to emit optical radiation after point A, when the RS weakens dramatically. This residual emission has a power-law light curve with a slope α ∼ 1. It is approximately described by the formula (see Beloborodov 2005b, § 7), α = p + 1 16 + 5p + 1 8γ , (3) where γ is the adiabatic index of the shocked ejecta. The RS is only mildly relativistic, and adiabatic cooling quickly reduces the sound speed to a non-relativistic value, so γ ≈ 5/3. Equation (3) then gives α ≈ 1.1 for p = 2.3. The RS recovery near point C makes an additional contribution to optical emission and produces the shoulder in the light curve. In contrast to the X-ray afterglow, it does not break at point C. We ran a number of explosion models with different Γej(Eej) and found that the optical shoulder usually accompanies the X-ray plateau. This feature is less prominent if Etail < Ehead (then it is buried by the residual optical emission from the head). For high ǫB ∼ 0.1, the optical-emitting electrons enter the fast-cooling regime before point C and the optical light curve has a plateau with a break at C, similar to that in the X-ray band. Figure 2f shows spectral indices βO and βX in the optical and X-ray (1 keV) bands for the RS emission. The X-ray break at the end of the plateau is not accompanied by any significant change of βX . This is consistent with observations (O'Brien et al. 2006). We ran the same model but with three bumps in external density n1(r) (Fig. 3). Fig- ure 3d compares the afterglow with and without the bumps. FS emission is barely changed (this conclusion is similar to that of Nakar & Granot 2006), and RS emission is changed significantly. The bump decelerates the blast, which weakens the FS and makes the RS stronger (γ43 temporarily increases; see Fig. 3b). The electron spectrum injected at the RS is then shifted to higher energies ∝ (γ43 − 1) and its synchrotron emission increases. The RS re-brightening is especially large if the bump occurs when the blast is near point C. The -- 7 -- mechanical model does not take into account sound waves or shock waves that may be gen- erated in the blast when it hits a bump, so the dynamics shown in Figure 3 is approximate. Nevertheless, this calculation well illustrates the difference between FS and RS. 3. DISCUSSION If the FS dominates afterglow production, several puzzles arise. We list these problems below and discuss how they are avoided/solved in the proposed RS model. 1. The FS model predicts much stronger emission at 102 − 103 s than observed. The FS does not stop producing the early afterglow -- it cannot weaken until Γ is reduced and tobs ∼ r/Γ2c becomes large -- while data suggest that the X-ray afterglow is temporarily suppressed at early times tobs ∼ 102 − 103 s. The RS weakens abruptly at tobs ∼ 102 s, when it crosses the head of the ejecta. This weakening explains the observed "lack" of early X-ray afterglow. 2. The FS model does not easily explain the shallow decay (plateau) in the X-ray light curve at tobs ∼ 103 − 104 s (Nousek et al. 2006; Zhang 2007; Granot 2006). Its predicted luminosity in the fast-cooling regime is L ∼ ǫeE/tobs, where E is the blast-wave energy. It can give a flat light curve if E(tobs) grows, i.e. if energy is gradually injected into the blast wave from the ejecta tail. However, this explanation invokes a huge energy injection (e.g. an increase of E by a factor of 3 does not help, cf. Fig. 2). A plateau of 1-2 orders in tobs would require a tail ∼ 100 times more energetic than the head of ejecta. On the other hand, the head should emit the prompt γ-ray burst that strongly dominates the observed energy output. This would require an extremely high radiative efficiency for the prompt emission. The RS produces the plateau BC as it recovers after the steep transition to the tail of the ejecta and enters the shallow part of the stratification function Γej(Eej) (Figs. 1 and 2a). In this scenario, Etail ∼ Ehead and 1 − 10% efficiency of the prompt emission is sufficient. 3. The observed peculiar break at the end of the plateau is difficult to reconcile with the FS model. The end of energy injection in the model causes a steepening of Γ(r), and the FS emission must respond to this steepening in all bands, both fast-cooling and slow-cooling. By contrast, the observed break often shows up only in X-rays (Panaitescu et al. 2006). Such chromatic breaks could appear when the X-ray spectrum steepens (νc passes through the XRT band), but this is excluded: the X-ray spectrum does not change across the break. This break is reproduced by the RS model (point C in Figs. 1-2), with no abrupt changes in βX. Emission in slow-cooling bands does not show a break; instead, a shoulder in the -- 8 -- Fig. 3. -- Same explosion model as in Fig. 2, but with three density bumps in the external density n1(r). Panel (a) shows the positions and amplitudes of the bumps. (b) Evolution of the RS relative Lorentz factor, γ43. (c) Evolution of pressures in the FS and RS, pf and pr. (d) Same as Fig. 2e, but for the model with bumps. For comparison, the model without bumps (from Fig. 2) is shown by thin curves in all panels. -- 9 -- light curve is predicted (§ 2). 4. To explain the second break that was observed in many afterglows at tobs ∼ 105 s, the FS model assumes a small beaming angle of the ejecta, θjet. When Γ decreases below θ−1 jet the "jet break" must inevitably appear in all bands (Rhoads 1999). By contrast, the observed breaks at ∼ 105 s are often chromatic or, in some bursts, not seen at all (e.g. Zhang 2007). The RS model predicts the second break when the shock reaches the end of ejecta or faces a steep decline in ρej (point D in Figs. 1-2). Its observed time is ∼ 1(Γend/20)−2 days. Afterglow steepens immediately only in fast-cooling bands.2 No dramatic change occurs in the blast deceleration at this point, and the slowly-cooling emission steepens with a delay. 5. The FS model does not easily explain re-brightenings observed in some optical afterglows. The response of FS emission to bumps in external density is weak (see Fig. 3 and Nakar & Granot 2006), while observed re-brightenings are strong, up to one order in flux. The RS emission is sensitive to external bumps (Fig. 3). Inhomogeneities in the ejecta may also cause bumps in the light curve. All five problems of the FS model ultimately have one common reason: changes in emission must invoke significant changes in Γ. By contrast, the RS model does not require strong deviations of Γ(r) from a power-law, and the afterglow light curves are shaped by the ejecta stratification. As a result, the temporal properties of the RS emission are consistent with observations, while those of the FS are not. As a solution, we propose that the entire afterglow is produced by the RS. The FS may be invisible for two reasons: (1) Magnetic fields are too weak in the external medium and not sufficiently amplified by the FS. If ǫB ∼ 10−7, the FS emission is negligible. (2) Electrons are not efficiently accelerated in the FS (e.g. diffusive acceleration is inefficient in ultra-relativistic shocks; see Beloborodov 2005b). By contrast, the ejecta may carry strong magnetic fields. Besides, the RS is mildly relativistic, which makes electron acceleration more plausible. This leaves the RS as the main producer of afterglow, and consistency with data is achieved. This model has interesting implications. The mean energy per electron in the RS is much lower than it would be in the FS model (while the number of electrons is much larger, and the dissipated energy in the RS is not much below that in the FS). Emission from 2 The light-curve slope becomes α = 2 + β if the RS weakens sharply at D -- the slope is then controlled by photons emitted at moment D and arriving with a delay because of the spherical curvature of the RS. By contrast, the usual jet model predicts α = p after the break. -- 10 -- most electrons is self-absorbed, and only a tail of their spectrum contributes to the X-ray and optical afterglow. The net radiative output of the blast wave peaks at the cooling frequency νc; however, an energetic fraction ∼ (ν/νc)(3−p)/2 is emitted at low ν ≪ νc. The RS afterglow should be very bright in the infrared bands, and this prediction may be tested by observations. If afterglow is indeed produced by the shocked ejecta rather than the shocked external medium, it carries important information about the explosion. We find that GRB ejecta are made of two blocks -- head and tail -- with different Γej and comparable kinetic energy. Its magnetic field is well below equipartition, ǫB ≪ 1. The RS emission is sensitive to the stratification function Γej(Eej), which explains the diversity of afterglows. We have shown here one example with a flat plateau, and other examples will be presented in an accompanying paper. The plateau has a larger slope α (and a longer duration) if Γej(Eej) is less concave at the beginning of the tail. The "steep decay + plateau" shape disappears if Γej decreases gradually after point A instead of jumping from ∼ 300 to < 100. In our models, the plateau often ends with a small flare, which is also observed (O'Brien et al. 2006). The study of radio emission expected in the RS model will be published elsewhere. Self- absorption limits the peak of radio afterglow, and observed light curves provide additional constraints on the model. The inverse Compton emission of high-energy γ-rays will also be investigated elsewhere. At least two high-energy components are expected: a brief flash that is produced through upscattering of prompt γ-rays by the electrons accelerated at the RS (Beloborodov 2005a) and the usual synchrotron self-Compton emission that can extend to longer times. This work was supported by NASA Swift grant, cycles 2 and 3. REFERENCES Beloborodov, A. M. 2005a, ApJ, 618, L13 Beloborodov, A. M. 2005b, ApJ, 627, 346 Beloborodov, A. M., & Uhm, Z. L. 2006, ApJ, 651, L1 (BU06) Granot, J. 2006, in SWIFT and GRBs: Unveiling the Relativistic Universe (astro-ph/0612516) M´esz´aros, P., & Rees, M. J. 1997, ApJ, 476, 232 -- 11 -- M´esz´aros, P., & Rees, M. J. 1999, MNRAS, 306, L39 Nakar, E., & Granot, J. 2006, preprint (astro-ph/0606011) Nousek, J. A., et al. 2006, ApJ, 642, 389 O'Brien, P., et al. 2006, ApJ, 647, 1213 Panaitescu, A., & Kumar, P. 2004, MNRAS, 350, 213 Panaitescu, A. et al. 2006, MNRAS, 369, 2059 Piran, T. 2004, Rev. Mod. Phys., 76, 1143 Rees, M. J., & M´esz´aros, P. 1998, ApJ, 496, L1 Rhoads, J. E. 1999, ApJ, 525, 737 Roming, P. W. A. et al. 2006, ApJ, 652, 1416 Sari, R., & Piran, T. 1999, ApJ, 517, L109 Sari, R., Piran, T., & Narayan, R. 1998, ApJ, 497, L17 Zhang, B. 2007, Adv. Space Rev., in press (astro-ph/0611774) This preprint was prepared with the AAS LATEX macros v5.2.
astro-ph/9807327
1
9807
1998-07-31T01:28:31
Determination of broadening functions using the Singular Value Decomposition (SVD) technique
[ "astro-ph" ]
Cross-correlation function (CCF) has become the standard tool for extraction of radial-velocity and broadening information from high resolution spectra. It permits integration of information which is common to many spectral lines into one function which is easy to calculate, visualize and interpret. However, CCF is not the best tool for many applications where it should be replaced by the proper broadening function (BF). Typical applications requiring use of the BF's rather than CCF's involve finding locations of star spots, studies of projected shapes of highly distorted stars such as contact binaries (as no assumptions can be made about BF symmetry or even continuity) and [Fe/H] metallicity determinations (good baselines and avoidance of negative lobes are essential). It is stressed that the CCF's are not broadening functions. The note concentrates on the advantages of determining the BF's through the process of linear inversion, preferably accomplished using the Singular Value Decomposition (SVD). Some basic examples of numerical operations are given in the IDL programming language.
astro-ph
astro-ph
Determination of broadening functions using the Singular Value Decomposition (SVD) technique Slavek Rucinski Canada-France-Hawaii Telescope Co. Kamuela, HI 96743, USA Abstract. Cross-correlation function (CCF) has become the standard tool for extraction of radial-velocity and broadening information from high resolution spectra. It permits integration of information which is common to many spectral lines into one function which is easy to cal- culate, visualize and interpret. However, CCF is not the best tool for many applications where it should be replaced by the proper broadening function (BF). Typical applications requiring use of the BF's rather than CCF's involve finding locations of star spots, studies of projected shapes of highly distorted stars such as contact binaries (as no assumptions can be made about BF symmetry or even continuity) and [Fe/H] metallicity determinations (good baselines and avoidance of negative lobes are essen- tial). It is stressed that the CCF's are not broadening functions. The note concentrates on the advantages of determining the BF's through the pro- cess of linear inversion, preferably accomplished using the Singular Value Decomposition (SVD). Some basic examples of numerical operations are given in the IDL programming language. 1. Convolution and cross-correlation Convolution is an operation that the nature does for us. We seldom see "naked" functions. These could be a convolution of a spectrum with the spectrograph's instrumental profile or with the radial component of the micro-turbulence ve- locity field in the stellar atmosphere or with a broadening function due to rapid rotation of a star. Thus, instead of a function f (u), we observe a function h(x) which is a convolution with some other broadening function (BF), g(x): h(x) = Z +∞ −∞ f (u) g(x − u) du = f (x) ∗ g(x) This natural process can be easily simulated in numerical packages (examples in the IDL programming language are marked by a command-line prompt IDL>) either by a special operator: IDL> h = convol(f, h) or through the Fourier-transform multiplication and its inverse: IDL> h = float(fft(fft(f,-1)*fft(g,-1),+1)) Cross-correlation is an operation which for real functions differs from the convolution really only in the symmetry of the arguments. For complex functions things are slightly different (real and imaginary parts have different symmetries), 1 but astronomers observe real spectra so we do not have to worry about the mathematical nuances. c(x) = Z +∞ −∞ f (u) g(u + x) du = f (x) ⋆ g(x) The cross-correlation (note the different asterisk above) function can be com- puted numerically using: IDL> lag = findgen(201) - 100 IDL> c = c correlate(f,g,lag) 2. Broadening functions Suppose we observe a sharp-line (S(λ)) and a broad-line (P (λ)) spectra and we want to determine the broadening and any other differences which make the latter spectrum more interesting than the former. The function B(λ) can be a rotational broadening function for a single, rapidly-rotating star, or a more complex profile for two components of a binary, or for a star with spots (where they would show as indentations in the function). The sharp-line spectrum S(λ) is not free of some broadening. This can be the thermal broadening of lines or micro-turbulence or some other mechanisms; we call them jointly T (λ). Thus, schematically, the sharp-line spectrum can be written as: S(λ) = (Xi aiδ(λi)) ∗ T (λ) while the broad-line spectrum, broadened additionally by B(λ), can be written as: P (λ) = S(λ) ∗ B(λ) = (Xi aiδ(λi)) ∗ T (λ) ∗ B(λ) The cross-correlation (CCF) with the sharp-line spectrum is frequently taken as an estimate of B(λ): C(λ) = S(λ) ⋆ P (λ) = S(λ) ⋆ (S(λ) ∗ B(λ)) = T (λ) ∗ T (λ) ∗ B(λ) The new function B(λ), is not identical to B(λ), because it inherits the common broadening components (such as thermal, micro-turbulence, instrumental) from both spectra. Tonry & Davis (1979) showed that if those common components are represented by Gaussians, the addition is quadratical, which for these func- tions means repeated convolutions. Thus, CCF cannot be really used to replace the broadening function. But it can give us some approximation of it and will remain a useful tool to have some preliminary estimate on the degree of the line broadening. For symmetrical broadening functions, it will remain the simplest tool to determine the radial velocities simultaneously from many spectral lines. The differences between the BF and the CCF can be seen when an artificial broadened spectrum is created by a convolution and then the resulting spectrum is subject to the CCF operation. The result (Figure 1) is obviously different from the BF: The CCF shows negative baseline excursions and, most worryingly, it shows the "peak-pulling" effect which would lead to an under-estimate of the individual component velocities. While this last problem can be overcome by applying the TODCOR technique (Zucker & Mazeh 1994), we clearly see that the CCF is not the BF. 2 Figure 1. The left panel shows a sharp-line spectrum (upper part) and the result of convolving it with the broadening function for a con- tact binary (lower part); this function is shown in the right panel by the dotted line (BF). The cross correlation of the broadened spectrum (with added noise to have S/N = 100) with the sharp spectrum is shown in the right panel by the continuous line. Notice the negative baseline and the peak-pulling in the CCF. 3. The Fourier transform de-convolution Some attempts to determine the broadening functions (Anderson et al. 1983) utilized the well known property of the Fourier transforms of a correspondence between convolutions and multiplications in the two relevant domains. Thus, a convolution: P (λ) = S(λ) ∗ B(λ) transformed with the Fourier transform F changes into a product of the trans- forms: F {P (λ)} = F {S(λ)} · F {B(λ)} Therefore, the broadening function can be restored with: B(λ) ≃ F −1 {F {P (λ)} /F {S(λ)}} This can be done in a compact way with: IDL> b = float(fft(fft(p,-1)/fft(s,-1),+1)) While the mathematical background is simple and easy, the practice is just the opposite. First, the resulting B(λ) spans the whole spectral window, so that one determines a lot of zeroes; there is no "compression" information whatsoever. But, more importantly, the division operation usually ... does not work: the high frequency noise becomes amplified and some sort of frequency filtering is needed. The result may then actually depend on the applied filter. Some authors (see for example the large collection of works of D. F. Grey) use spectra transformed into the frequency domain without the division step. This can be done for broadening mechanisms describable by simple functions or obeying some symmetries, but fails in cases of spots or of BF's of close binary systems. 3 4. Convolution in the formalism of linear equations There are two main issues that re-casting convolution into a set of linear equa- tions can resolve. These are: (1) How to channel information over the whole spectrum (say 2000 pixel long) into the BF window (say 200 pixels long)? (2) How to utilize all information contained in sharp-line spectra and remove the influence of the noise in the continuum (which carries no information)? The convolution can be written as an over-determined system of linear equations which link a sharp-line spectrum ~S(n), via the broadening function ~B(m), with the broadened spectrum ~P (m). The mapping is through the "design matrix" dDes(m, n) which is formed from the sharp line spectrum ~S(n) by con- secutive vertical shifts by one element. In IDL, this can be done with a simple routine: function map4,s,m ; m - must be odd, n must be even n = n elements(s) & t = fltarr(m) # fltarr(n-m+1) ; t(m,n-m) = t(small,large-small) dimensions for j = 0,m-1 do for i = m/2,n-m/2-1 do t(j,i-m/2)=s(i-j+m/2) return,t end An example of using this routine to create a design matrix for a 201-pixel long window would be: IDL> des = map4(s,201). The program spectra must accordingly be trimmed to n-m+1 with: IDL> p = p(m/2:n-m/2-1). The sys- tem of equations has a familiar form of the over-determined linear set: ~B = ~P dDes 5. Singular value decomposition (SVD) One of the traditional ways of solving the system of equations above would be to transform it into a system of the "normal" equations of the size reduced from m × (n − m) to m × m by multiplication of both sides by the transpose of the design matrix. The result would be the BF defined in the least-squares sense, and it is possible to stop at this point. However, the Singular Value Decomposition technique also gives us such an answer, but -- in addition -- makes it possible to remove the influence of the continuum and its noise. The SVD technique is beautifully described in the "numerical techniques Bible" of Press et al. (1986). They present it as a somewhat magic black box 4 and for most users it is just fine. If you want to learn how the technique really works, then the books of Golub & Van Loan (1989) or Craig & Brown (1986) are probably the best references. The essence of the SVD is the property that one can represent any matrix by a product of 3 matrices; in our case: dDes = bU cW bV T . These matrices are: the column ortho-normal bU and bV and the diagonal matrix cW (this is really a vector containing the diagonal elements). The property of the columns in bU and bV T is that the following products, bU T bU = bI and bV T bV = bI, give the unity array bI (1 on the diagonal). dDes = bU cW bV T In IDL, the operation is represented by: IDL> svdc,des,w,u,v,/double. Here, des is the only input quantity and the remaining parameters are what the routine produces as output. The keyword /double is for higher precision and is optional. One can check the correctness of the operations by the following commands: IDL> wf = fltarr(m,m) IDL> for i = 0,m-1 do wf(i,i) = w(i) IDL> des check = u ## wf ## transpose(v) The three new matrices are all invertible: bU and bV are ortho-normal arrays, so that their inverses are just transposes, while the diagonal array cW is replaced by a diagonal array dW 1, with the diagonal elements containing the inverses, w1i,i = 1/wi: IDL> w1 = fltarr(m,m) IDL> for i = 0,m-1 do w1(i,i) = 1./w(i) The solution is given by: ~B = bVdW1(bU T ~P ) or schematically: ~B = bV dW 1 bU T ~P 5 Figure 2. The left panel shows the run of the singular values for the sharp-line spectrum in Figure 1. The right panel gives the solutions for the broad-line spectrum in the same figure (with added noise at S/N = 100) utilizing the first 5, 10, 20 and 30 singular values. Numerically: IDL> b = reform(v##w1##(transpose(u)##p)). A routine makes this simpler: IDL> b = svsol(u,w,v,p,/double). The elements of ~B are all independent, so that any -- even strange or discontinuous -- broadening functions can be restored as no condition imposed on the smoothness or sym- metry of the result. Note that, if only one sharp-line template is used, the decomposition operation svdc is done only once, for possibly many broad-line spectra p, each giving a separate solution b. 6. Advantages and disadvantages of the SVD approach On the positive side: (1) The problem can be treated as a linear-equations. (2) An "inverse" of the rectangular array dDes is possible. (3) The solution of ~B is defined in the least-squares sense (shortest modulus). (4) The result is the real broadening function. But there are also minuses: (5) One must solve a large system of, say, 2000 equations for 200 unknowns. (6) One must know a priori how many unknowns. (7) Initially, the results may turn out quite poor, because of the presence of plenty of linearly-dependent equations in the system (parts of spectra where the featureless continuum provides no broadening information). The SVD approach offers a simple resolution of (7) as it permits removal of the effects of the continuum in an objective way. The key element here are the singular values contained in the diagonal of cW . Since the solution involves 1/wi, small values in wi spoil the solution. These are exactly those problematic values that one wants to avoid. Thus, by rejecting of small values of wi, one can (i) remove the linearly dependent equations, (ii) diminish the influence of the noise from the continuum, (iii) reduce the influence of the computer round-off errors (which enter multiplied by the order of the problem) and (iv) reduce the number of the unknowns (because the system is usually not over-determined at all). All these properties are related to the "conditioning" of the array dDes. The reader is directed to the source texts on this subject for further reading. The important factor is max(wi)/min(wi) which provides an estimate on how many of the singular should be used. In practice, one can keep on adding 6 more wi and see the successively better solutions. The diagonal arrays cW and dW 1 will have then elements: wi = w0, w1, w2, wk, ..., wm−1 and w1i = 1/w0, 1/w1, 1/wk−1, 0, ..., 0 with k (we call it the order of solution) spanning the whole range 0 to m − 1. In IDL, this can be done by forming a square matrix of solutions: b = fltarr(m,m) for i = 0,m-1 do begin wb = fltarr(m) wb(0:i) = w(0:i) ; first i+1 singular values used, rest zero b(*,i) = svsol(u,wb,v,p,/double) end for as they contain the basis vectors in the spaces of the spectra and broadening functions, respectively. One can see this by analyzing a diagonalized system: We note in passing, that the arrays bU and bV have very special properties cW ~Z = ~D, obtained by keeping the same cW , with ~Z = bV T ~B and ~D = bU T ~P . practice, the diagonal of cW is a vector, so that zi = di/wi. Plotting the columns of bU and bV can tell one a lot about the conditioning of the solution. The solution of the diagonalized system would be then: ~Z = ~D/ ~W , but, in 7. A few notes on the SVD solutions The first question is: How far in k should one go and where to stop? The essential operation is to plot (usually in log units) the vector ~W (Figure 2). There are 3 parts of it: (1) the good, large singular values, (2) the small values usually representing the noise in the spectrum S(λ) and (3) the numerical errors. You may want to stop no later than at the kink below the good part. But the real "quality control" is the fit to ~P . If the error of the fit stops decreasing, you have found the right point (Figure 3). Beyond that point, you will start fitting the noise! However, it is useful to analyze the solutions for different orders k and see how they first improve and then get worse. Sometimes the fit will remain poor, in spite of the leveling of the error curve; this usually means a wrong choice of the sharp-line spectrum. The standard error of the fit can be calculated from: sig = fltarr(m) ; error pred = des ## transpose(b) ; predicted fits for i=0,m-1 do sig(i) = sqrt(total((pred(i,*)-p)^2)/m) Usually, even very low order solutions are well defined (Figure 2), but stop- ping early is not always advisable because this leads to a loss of resolution, as then not all basis vectors contribute. The solution which -- in principle -- has all elements in ~B independent suddenly acquires inter-element correlations. Thus, it may be advantageous to go to a highest possible order (k = m) and thus insure that elements of ~B are uncorrelated, and then decrease the noise by smoothing. One should be aware that the errors may be under-estimated for the case of truncated (k < m) solutions. While the prescriptions of Rix & White (1992) and Rucinski, Lu & Shi (1993) are based on the theory of the full SVD, the error analysis for the truncated case has not yet been done. In this situation, it may be advantageous to utilize techniques of the external estimates, such as the bootstrap or Monte Carlo. This subject certainly requires more work ... 7 Figure 3. The left panel shows the mean standard error of the fit for our example solutions utilizing progressively more singular values. It is obvious that the first 10 -- 15 singular values give an adequate fit (compare with Figure 2). The right panel shows the solution utilizing all 200 singular values, convolved with a Gaussian. This approach may offer a better control over the final resolution of the BF. 8. Conclusions One can position the SVD technique of linear broadening function restoration as located between the cross-correlation and the direct modeling of the spectra. The BF's determined with it are much better defined than the CCF's, and are true broadening functions, not their proxies. They also integrate the geometrical information from a spectrum, but give well defined baselines without the CCF's negative fringes. They do require a bit more computer work, but several numer- ical packages can easily handle large linear systems of equations involved here. Obviously, they cannot replace the spectrum synthesis, if these are needed, but can be a useful tool in their preparation. References Anderson, L., Stanford, D. & Leininger, D. 1983, ApJ, 270, 200 Craig, I,J.D. & Brown, J.C. 1986, Inverse Problems in Astronomy, (Bristol and Boston: Adam Hilger Ltd) Golub, G.H. & Van Loan, C.F. 1989, Matrix Computations, 2nd ed. (Baltimore: Johns Hopkins Univ. Press) Press, W.H., Teukolsky, S.A., Vetterling, W.T. & Flannery, B.P. 1986, Numeri- cal Recipes, Cambridge University Press (all editions) Rix, H.-W. & White, S.D.M. 1992, MNRAS, 254, 384 Rucinski, S.M. 1992, AJ, 104, 1968 Rucinski, S.M., Lu, W.-X. & Shi, J. 1993, AJ, 106, 1174; Tonry, J. & Davis, M. 1979, AJ, 84, 1511 Zucker, S. & Mazeh, T. 1994, ApJ, 420, 806 8
astro-ph/0207369
1
0207
2002-07-18T09:19:48
Gravitational Potential Reconstruction from Peculiar Velocity and Weak Lensing Measurements
[ "astro-ph" ]
We present an analytic method for rapidly forecasting the accuracy of gravitational potential reconstruction possible from measurement of radial peculiar velocities of every galaxy cluster with M > M_th in solid angle \theta^2 and over redshift range z_min < z < z_max. These radial velocities can be determined from measurement of the kinetic and thermal Sunyaev-Zeldovich effects. For a shallow survey with 0.2 < z < 0.4, coincident with the SDSS photometric survey, one mode of the gravitational potential (on length scales > 60 Mpc) can be reconstructed for every ~8 cluster velocity determinations. Deeper surveys require measurement of more clusters per S/N > 1 mode. Accuracy is limited by the ``undersampling noise'' due to our non-observation of the large fraction of mass that is not in galaxy clusters. Determining the gravitational potential will allow for detailed study of the relationship between galaxies and their surrounding large-scale density fields over a wide range of redshifts, and test the gravitational instability paradigm on very large scales. Observation of weak lensing by large-scale structure provides complementary information since lensing is sensitive to the tangential modes that do not affect the velocity.
astro-ph
astro-ph
To be submitted to ApJ Letters Preprint typeset using LATEX style emulateapj v. 16/07/00 2 0 0 2 l u J 8 1 1 v 9 6 3 7 0 2 0 / h p - o r t s a : v i X r a GRAVITATIONAL POTENTIAL RECONSTRUCTION FROM PECULIAR VELOCITY AND WEAK LENSING MEASUREMENTS Olivier Dor´e1, Lloyd Knox2 and Alan Peel2 1 Institut d'Astrophysique de Paris, 98bis boulevard Arago, 75014 Paris, FRANCE, email: [email protected] 2 Department of Physics, University of California, Davis, CA 95616, USA, email: [email protected] To be submitted to ApJ Letters ABSTRACT We present an analytic method for rapidly forecasting the accuracy of gravitational potential recon- struction possible from measurement of radial peculiar velocities of every galaxy cluster with M > Mth in solid angle θ2 and over redshift range zmin < z < zmax. These radial velocities can be determined from measurement of the kinetic and thermal Sunyaev–Zeldovich effects. For a shallow survey with 0.2 < z < 0.4, coincident with the SDSS photometric survey, one mode of the gravitational potential (on length scales & 60 Mpc) can be reconstructed for every ∼ 8 cluster velocity determinations. Deeper surveys require measurement of more clusters per S/N > 1 mode. Accuracy is limited by the "under- sampling noise" due to our non–observation of the large fraction of mass that is not in galaxy clusters. Determining the gravitational potential will allow for detailed study of the relationship between galaxies and their surrounding large–scale density fields over a wide range of redshifts, and test the gravitational instability paradigm on very large scales. Observation of weak lensing by large–scale structure provides complementary information since lensing is sensitive to the tangential modes that do not affect the velocity. Subject headings: cosmology: theory – cosmology: observation – cosmology: weak lensing – cosmology: peculiar velocities – galaxies: formation – galaxies: evolution 1. introduction Although there are a variety of techniques for measuring the statistical properties of cosmological density fluctua- tions, we know of only three types of observations from which the density field itself may be reconstructed: weak lensing (Mellier, 1999; Bartelmann & Schneider, 2001), pe- culiar velocities (Strauss & Willick, 1995) and galaxy rota- tion vectors (Lee & Pen, 2000). A map of the density field, combined with galaxy surveys, would be a highly valuable aid to understanding the formation of galaxies and clus- ters of galaxies. Numerical simulations could be performed with realizations of initial conditions constrained by our knowledge of the density field, allowing object–by–object comparison between theory and observation, rather than solely statistical comparison. A map could also provide a guide to observers who may, for example, wish to search for the luminous tracers of the filamentary density struc- tures. Reconstruction of the density field from galaxy peculiar velocities was pioneered by Dekel et al. (1990). The radial component of peculiar velocities of galaxies can be deter- mined by inferring the distance and then subtracting off the Hubble flow contribution to the redshift. Galaxy pecu- liar velocity determinations, because they rely on distance determinations, are only useful at z . 0.1. Peculiar velocities of galaxy clusters determined from observations of the Sunyaev–Zeldovich (SZ) effects (Sun- yaev & Zeldovich, 1980) do not rely on a distance determi- nation. The peculiar velocity signal arises from the clus- ter's radial motion with respect to the cosmic microwave background (CMB). High–resolution (. 1′), multi–frequency observations of galaxy clusters can be used to determine the radial velocities of galaxy clusters with an accuracy that is independent of the distance to the galaxy cluster1. 1This independence only fails for clusters at z . 0.2 which suffer 1 Errors as low as 100 km s−1 may be achievable (Holder, 2002). In addition to potential reconstruction, velocity measurements will be useful for cosmological parameter estimation (Peel & Knox, 2002). Here we present a method for rapidly forecasting the accuracy of density field reconstruction from peculiar ve- locity measurements and apply it to surveys with various redshift ranges. We also show how weak lensing obser- vations can provide complementary information (Mellier, 1999; Bartelmann & Schneider, 2001). Our analytic method for forecasting results assumes a uniform and continuous velocity field map. Even if this map were derived from noiseless peculiar velocity mea- surements, there would still be an effective noise contribu- tion from the fact that most of the mass in the Universe is not in galaxy clusters. Fortunately, as we quantify be- low, the contribution to the peculiar velocity variance from each logarithmic interval in wavenumber k drops as k−1 for relevant scales, so this noise from under–sampling is not overwhelmingly large. We consider three different survey types labeled "SDSS", "DEEP/VIRMOS" and "SZ" with redshift ranges 0.2 < z < 0.4, 0.7 < z < 1.4 and 0.2 < z < 2 respectively. Galaxy cluster redshifts are required to convert SZ mea- surements into gravitational potential maps. Thus, the redshift ranges of two survey types are subsets of those of optical surveys SDSS (Sloan Digital Sky Survey), DEEP II (Deep Extragalactic Evolutionary Probe; Davis et al. (2001)) and VIRMOS (Visible-Infrared Multi-Object Spec- trographs; Le F`evre et al. (2001)). The peculiar velocity surveys we imagine could be realized by targeted multi– frequency SZ follow–up on galaxy clusters identified in those optical surveys. The ambitious, deep "SZ" survey significant confusion with primary CMB anisotropy due to their large angular sizes. 2 Gravitational potential from peculiar velocities and weak lensing measurements is modelled after proposed SZ surveys which can locate galaxy clusters over a large z range. Again, the velocities can be determined from multi–frequency follow up (if nec- essary). The redshift determinations for this last survey type are left as a currently unsolved challenge. Radial peculiar velocities are directly sensitive to ra- dial gradients in the gravitational potential. Gravitational lensing is caused by tangential gradients in the gravita- tional potential; thus, in principle, lensing provides com- plementary information for potential reconstruction. How- ever, we show below that lensing is actually unlikely to add much to the density field reconstruction. The rea- son is simple: lensing observations are two–dimensional, whereas velocity observations are three–dimensional. The conclusion may be different if one considers lensing of many source populations with differing redshift distribu- tions (Hu & Keeton, 2002). 2. forecasting errors To begin, we assume that we have a continuous and uniform map of the velocity field in a box of volume V . The map is not noise–free, but has white noise with finite weight–per–comoving volume w; i.e, the variance of the error on the average velocity in some sub-volume V1 is σ2 V1 = 1/(wV1). For specificity we set our fiducial model to be h = 1 − Ωm = ΩΛ = 0.7 and σ8 = 1. For the moment we ignore the fact we can only make measurements along our past light cone and assume the map is of the velocity field today. We remove this oversim- plification below, but for now this idealization maintains homogeneity and allows a completely analytic analysis of the potential reconstruction errors in Fourier space. With the assumption of a potential flow (Peebles, 1993) ~v~k = i ~k H0 Φ~k(t0)x(t) x(t) ≡ 2 3 dD da (t) E(t)a2(t) Ωm (1) (2) where and E2(t) ≡ H 2(t)/H 2 (3) The noise and signal contributions to the variance of Φ~k are diagonal with diagonal entries given by 0 = Ωm/a3(t) + ΩΛ. N (k) = (cid:18) H0 kx0(cid:19)2 /w and S(k) = PΦ(k) = 9 k (cid:19)4 4 (cid:18) H0 Ω2 m 2π2 k3 ∆2(k) (4) (5) respectively where ∆2(k) ≡ k3Pδ(k)/(2π2) and Pδ(k) is the matter power spectrum. The square of the signal–to–noise ratio for the gravita- tional potential map is thus S(k)/N (k) = 1(cid:18) w wf (cid:19)(cid:18) ∆2(k) 0.21 (cid:19)2(cid:18) f ≡ (cid:0)100 km s−1(cid:1)2 ×(cid:18) Ωmh w−1 0.8 (cid:19)(cid:16) x0 1.1(cid:17)2 0.1 Mpc−1(cid:19)−5 k (6) where x0 ≡ x(t0) = 1.1 and ∆2(0.1 Mpc−1) = 0.8 for our fiducial model. We therefore expect measurements with (64 Mpc)3 S/N > 1 for every mode with 2π/L & k & 0.1 Mpc−1 where L3 = V . Further, there are many of these modes for a box of size L; the number in Fourier–space volume . (∆k)3 is 4 × 103(cid:0)(k/0.1 Mpc−1)(L/1000 Mpc)(∆k/k)(cid:1)3 We now address our value of wf above. A fundamen- tal limit on w is set by the small fraction of mass in galaxy clusters. The comoving number density of galaxy clusters with lower mass threshold 1014h−1M⊙ at z = 1 is about one per (64 Mpc)3 volume (Holder et al., 2001) which roughly corresponds to the volume of a 40 Mpc ra- dius sphere. However, a sphere containing this limiting mass in a uniform background only has a comoving ra- dius of about R = 9 Mpc. Estimating the velocity field averaged over R = 40 Mpc from just one sample of the velocity field measured over 9 Mpc introduces an under– sampling error with variance (at z = 1) of h(v40 − v9)2i = (117 km s−1)2. Thus we cannot do better than w−1 = (117 km s−1)2(64 Mpc)3 at z = 1. The dashed line in Fig. 1 shows the redshift dependence of this under–sampling noise which is due to evolution in ¯n. According to Holder (2002) and Nagai et al. (2002) there is also a fundamental limit on how well one can infer the peculiar velocity of a galaxy cluster from the (highly non–uniform) velocities of the gas, of about 100 km s−1. The result of adding these errors in quadrature is the solid line of Fig. 1. Wiener filtering can reduce the error, though we do not employ it here. The angular resolution, sensitivity and frequency cov- erage required to achieve a velocity measurement with a given precision have not yet been worked out in detail. From prior work (Haehnelt & Tegmark, 1996; Holzapfel et al., 1997; Kashlinsky & Atrio-Barandela, 2000; Aghanim et al., 2001), which ignores the velocity substructure of the gas, we find measurements with errors of 8µK translate into σv ≃ 100 km s−1, well–matched to the Holder/Nagai limit. A multi–mode detector on a large telescope with sensitivity of 100µKs1/2 could achieve 8µK on 20 clusters in 1 hour, or 16,000 in a month. We now consider observations of the radial component Fig. 1.- Inverse weight–per–solid angle, w−1, from under– sampling noise (dashed line) and from under–sampling noise plus a 100 km s−1 error on each galaxy cluster added in quadrature (solid line). The dotted line is R(z)/30 Mpc where 4/3πR3(z)¯n(z) = 1 defines R(z) and ¯n(z) is the comoving average number density of clusters. Dor´e et al. 3 of peculiar velocities on our past light cone. The isotropic but inhomogeneous geometry suggests the mode decom- position Φ(γ, r) = Xlm Z ∞ 0 dkk2 Φlm(k)jl(kr)Ylm(γ) (7) as used by e.g. Stebbins (1996). The line–of–sight radial velocity generated by this potential is vr(γ, r) = ix(r)/H0Xlm Z ∞ 0 dkk3 Φlm(k)j ′ l(kr)Ylm(γ) (8) where we have changed our independent variable from t to conformal distance, r and j ′ l(y) ≡ d/dy jl(y). Note that all the time–dependence on the right–hand side is carried by x(r), thus we are discussing reconstruction of the grav- itational potential on our past light cone extrapolated by linear theory to what it would be today. Our discretized model of the data is vi = Xblm Av(i; b, l, m) Φblm + ni (9) where n is the error (whose statistical properties were just discussed above), Av is implicitly defined by Eq. 8 and b enumerates the k bins: 1 dk Φlm(k). (10) Φblm ≡ ∆kb Z kb+∆kb/2 kb−∆kb/2 The minimum–variance estimate of Φblm is Φblm = Xb′l′m′ii′ W −1 vΦ (blm, b′l′m′) × Av(i; b′, l′, m′)N −1 ii′ vi′ where Nij = hninji, W −1 vΦ (blm, b′l′m′) = hδΦblmδΦb′l′m′i is the noise covariance matrix of Φblm given by WvΦ(blm, b′l′m′) = Xii′ and δ Φblm ≡ Φblm − Φblm. Av(i; b, l, m)N −1 ii′ A∗ (11) (12) (13) v(i′; b′, l′, m′) where Il(krs, θs) = Z rs 0 dr rrs (rs − r) jl(kr) D(η0 − r) a(η0 − r) W (krθs) (16) (Stebbins, 1996), and W (krθs) = 2J1(krθs)/(krθs) (Jain & Seljak, 1997) incorporates the fact that we smooth the signal over radius θs = 2.5′. Again, this is block–diagonal in l and m. However, the structure in the radial wave– number k is highly degenerate. For each l, m sub–matrix, the k, k′ dependence can be written as an outer product of a single vector. Such matrices have only one non–zero eigenvalue, a manifestation of the inability of lensing mea- surements to distinguish between different radial modes. 3. results In Figure 2 we compare signal and noise for ∆k = k for several values of l assuming w(z) as shown in Fig. 1. One can see that S/N > 1 measurements are possible on large spatial and angular scales. The sharp low k cutoffs are an effect of the finite size of the surveys: kmin = l/L where L is the radial depth of the survey since jl(kr) ≃ 0 for k < l/r. The velocity results can be understood quantitatively by multiplying S/N from Eq. 6 by the number of modes in a bin of width ∆k: ∆kL/(2π). Now we turn to the weak lensing panel of Fig. 2. We assume a noise level of w = 9 × 109 which can be achieved with a galaxy number density of n = 30 gal. arcmin−2 and intinsic galaxy ellipticity of σǫ = 0.2 (Kaiser, 1998; van Waerbeke et al., 1999). We further assume a negli- gible width to the distribution of these galaxies centered at zs = 1. The results can be understood analytically from Eq. 15 and the approximation (Stebbins, 1996) Il = pπ/(2l)(1/l − 1/(krs)) for krs > l and Il = 0 otherwise. Unlike with velocity the higher l values are reconstructed with less noise due to the sensitivity of the convergence to tangential gravitational potential variations. Above we have plotted k6 b /WvΦ(blm, blm) as an indica- tor of the variance expected on a measurement of k3 b Φblm. However, the variance of the error on this mode is actu- With the uniform sampling assumption, Wv Φ for recon- structing Φblm is given by ∆kb′ H0 b k3 b′ δll′ δmm′k3 l(kb′ r). Wv Φ(blm, b′l′m′) = ∆kb ×R drr2x2(r)w(r)j ′ H0 l (kbr)j ′ (14) The matrix is block–diagonal with zero entries for any el- ements with l 6= l′ or m 6= m′. As one expects from statistical isotropy Wv Φ does not depend on m. The signal matrix is diagonal, though the effect of bin- ning the uncorrelated k modes is to reduce the variance by ∆k: h Φblm Φblmi = P Φ(k)/∆k. In the results section we compare signal and noise by plotting k5∆k/Wv Φ where the k5 factor makes a dimensionless quantity and the ∆k makes the corresponding signal quantity, k5P Φ(k) = (4π)2 ×k3PΦ(k), independent of ∆k. The weight matrix for potential reconstruction from a weak lensing convergence map with uniform weight per solid angle w: Wκ Φ(blm, b′l′m′) = δll′ δmm′ wl2(l + 1)2 ×∆kbk2 b′ Il(kbrs, θs)Il(kb′ rs, θs) b ∆kb′ k2 (15) Fig. 2.- Dimensionless diagonal noise variance (∆kk5/WX Φ) for binned modes of the gravitational potential with bin widths ∆k = k for several values of l, for our three velocity surveys (X = v) and the lensing survey (X = κ). The expected signal variance, k5PΦ(k) is also shown for comparison. 4 Gravitational potential from peculiar velocities and weak lensing measurements b W −1 (e.g., Bond (1995)). P Φ/∆kW 1/2 v Φ ally given by k6 v Φ(blm, blm). This inverse is not well– defined because of eigenmodes with zero eigevnalues. In Fig. 3 we plot the eigenvalues of the signal–to–noise matrix W 1/2 v Φ The total weight from a combined weak lensing and pe- culiar velocity survey is simply given by adding the indi- vidual weights. The eigenvalue spectrum for this combina- tion is identical to the peculiar velocity survey eigenvalue spectrum, except for the increased value of one eigenvalue per l, m mode. 4. discussion The largest l value of modes with S/N > 1, lmax, entails a minimum survey size of θmin = 5◦(80/lmax). From Fig. 3 we see lmax ≃ 80 and therefore θmin ∼ 5◦ for the two deeper surveys and lmax ≃ 35 and therefore θmin ≃ 10◦ for SDSS. There are many advantages to the shallower SDSS sur- vey. The spatial number density of clusters is higher, re- ducing the under–sampling noise. Despite the increased value of θmin, the clusters in solid angle θ2 min are actually fewer for the shallow survey (358) than for the deeper sur- veys (500 and 1000). This is an advantage since at fixed detector sensitivity the requisite observing time is propor- tional to the number of clusters, not the survey solid angle. Nearer clusters are easier to observe in the optical and in X–rays, so redshift and temperature determinations will take less time. Finally, a nearby map would be easier to compare to other tracers of the density field than a more distant map would be. Increasing the survey solid angle θ2 not only reduces lmin, but also increases spectral resolution (i.e., decreases δl = 2π/θ = 12(30◦/θ)) and increases the number of an- gular modes with S/N > 1: Nθ = θ2 4π lmax Xl=lmin (2l + 1) = 27(cid:18) θ 35 (cid:19)2 30◦(cid:19)2(cid:18) lmax − π. (17) Thus for a θ = 30◦ survey with 0.2 < z < 0.4 the total number of modes with S/N & 1 is roughly Nθ × 15 ≃ 360. Covering the entire π steradians of the SDSS survey would require measuring 36,000 clusters and would deliver ∼ 4, 500 modes. 5. conclusions We have shown that galaxy cluster peculiar velocities can be used to make high signal–to–noise maps of the gravitational potential, and therefore matter density, on very large scales. A limiting source of uncertainty in these maps is what we have called the under–sampling noise due to the fact that only a small fraction of mass is in clusters. The under–sampling noise can be reduced by lowering the mass threshold but the lower the mass of the cluster, the harder the measurement of its peculiar velocity. Much more work is needed to understand the demand on experi- mental resources and to optimize observing strategies. We have made several arguments that favor shallower surveys over deeper ones. The Φ and density maps will have numerous applica- tions. Cross–correlations with various tracers of the den- sity field (red galaxies, blue galaxies, infrared galaxies, lensing of galaxies, lensing of the CMB, integrated Sachs– Wolfe effect on CMB, ...) will be of great interest. The greatest drawback of the maps is their inability to probe scales below the typical separation between clusters. We thank G. Holder for his encouragement and for shar- ing the ¯n(z) calculation used in Holder et al. (2001) and L. van Waerbeke and S. Colombi for several useful conversa- tions. O.D. acknowledges financial support from "Soci´et´e de Secours des Amis des Sciences". REFERENCES Aghanim, N., G´orski, K. M., & Puget, J.-L. 2001, A&A, 374, 1 Bartelmann, M. & Schneider, P. 2001, Phys. Rep., 340, 291 Bond, J. R. 1995, Physical Review Letters, 74, 4369 Davis, M., Newman, J. A., Faber, S. M., & Phillips, A. C. 2001, in Deep Fields, 241+ Dekel, A., Bertschinger, E., & Faber, S. M. 1990, ApJ, 364, 349 Haehnelt, M. G. & Tegmark, M. 1996, MNRAS, 279, 545+ Holder, G. 2002, in preparation Holder, G., Haiman, Z. ., & Mohr, J. J. 2001, ApJ, 560, L111 Holzapfel, W. L., Ade, P. A. R., Church, S. E., Mauskopf, P. D., Rephaeli, Y., Wilbanks, T. M., & Lange, A. E. 1997, ApJ, 481, 35+ Hu, W. & Keeton, C. 2002, astro-ph/0205412 Jain, B. & Seljak, U. 1997, ApJ, 484, 560+ Kaiser, N. 1998, ApJ, 498, 26+ Kashlinsky, A. & Atrio-Barandela, F. 2000, ApJ, 536, L67 Le F`evre, O., Vettolani, G., Maccagni, D., Mancini, D., Mazure, A., Mellier, Y., Picat, J. P., Arnaboldi, M., Bardelli, S., Bertin, E., Busarello, G., Cappi, A., Charlot, S., Colombi, S., Garilli, B., Guzzo, L., Iovino, A., Le Brun, V., Mathez, G., Merluzzi, P., McCracken, H. J., Pell`o, R., Pozzetti, L., Radovich, M., Ripepi, V., Saracco, P., Scaramella, R., Scodeggio, M., Tresse, L., Zamorani, G., & Zucca, E. 2001, in Deep Fields, 236+ Fig. 3.- Signal–to–noise eigenvalue spectra for individual angular modes. The boxes show the single non–zero eigenvalues of the weak lensing survey. The SDSS panel includes additional curves for l = 10 (top curve) and l = 50. Lee, J. & Pen, U. 2000, ApJ, 532, L5 Mellier, Y. 1999, ARA&A, 37, 127 Nagai, D., Kravtsov, V., & Kosowsky, A. 2002, in preparation Peebles, P. J. E. 1993, Principles of physical cosmology (Princeton Series in Physics, Princeton, NJ: Princeton University Press, -c1993) Peel, A. & Knox, L. 2002, in Proceedings of the 5th International UCLA Symposium on Sources and Detection of Dark Matter and Dark Energy in the Universe, 20-22 February 2002 Stebbins, A. 1996, astro-ph/9609149 Strauss, M. A. & Willick, J. A. 1995, Phys. Rep., 261, 271 Sunyaev, R. A. & Zeldovich, I. B. 1980, MNRAS, 190, 413 van Waerbeke, L., Bernardeau, F., & Mellier, Y. 1999, A&A, 342, 15
astro-ph/0208002
2
0208
2002-10-07T04:25:54
Importance of Cluster Structural Evolution in Using X-ray and SZE Galaxy Cluster Surveys to Study Dark Energy
[ "astro-ph" ]
We examine the prospects for measuring the dark energy equation of state parameter w within the context of the still uncertain redshift evolution of galaxy cluster structure. We show that for a particular X-ray survey (SZE survey) the constraints on w degrade by roughly a factor of 3 (factor of 2) when one accounts for the possibility of non--standard cluster evolution. With followup measurements of a cosmology independent, mass--like quantity it is possible to measure cluster evolution, improving constraints on cosmological parameters (like w Omega_M). We examine scenarios where 1%, 10% and 100% of detected clusters are followed up, showing that even a modest followup program can enhance the final cosmological constraints. For the case of followup measurements on 1% of the cluster sample with an uncertainty of 30% on individual cluster mass--like quantities, constraints on w are improved by a factor of 2 to 3. For the best case scenario of a zero curvature universe, these particular X-ray and SZE surveys can deliver uncertainties on w of ~4% to 6%.
astro-ph
astro-ph
submitted to ApJ June 14, 2002 Preprint typeset using LATEX style emulateapj v. 26/01/00 2 0 0 2 t c O 7 2 v 2 0 0 8 0 2 0 / h p - o r t s a : v i X r a IMPORTANCE OF CLUSTER STRUCTURAL EVOLUTION IN USING X -- RAY AND SZE GALAXY CLUSTER SURVEYS TO STUDY DARK ENERGY Subhabrata Majumdar1 & Joseph J Mohr1,2 submitted to ApJ June 14, 2002 ABSTRACT We examine the prospects for measuring the dark energy equation of state parameter w within the context of the still uncertain redshift evolution of galaxy cluster structure. We show that for a particular X -- ray survey (SZE survey) the constraints on w degrade by roughly a factor of 3 (factor of 2) when one accounts for the possibility of non -- standard cluster evolution. With followup measurements of a cosmology independent, mass -- like quantity it is possible to measure cluster evolution, improving constraints on cosmological parameters (like w & ΩM ). We examine scenarios where 1%, 10% and 100% of detected clusters are followed up, showing that even a modest followup program can enhance the final cosmological constraints. For the case of followup measurements on 1% of the cluster sample with an uncertainty of 30% on individual cluster mass -- like quantities, constraints on w are improved by a factor of 2 to 3. For the best case scenario of a zero curvature universe, these particular X -- ray and SZE surveys can deliver uncertainties on w of ∼4% to 6%. Subject headings: cosmic microwave background -- galaxies: clusters -- cosmology: theory 1. INTRODUCTION Galaxy clusters have been used extensively to determine the cosmological matter density parameter and the ampli- tude of density fluctuations. Cluster surveys in the local universe are particularly useful for constraining a combi- nation of the matter density parameter ΩM and the nor- malization of the power spectrum of density fluctuations (we describe the normalization using σ8, the rms fluctua- tions of overdensity within spheres of 8h−1 Mpc radius; i.e. Henry 1997; Viana & Liddle 1999; Reiprich & Bohringer 2002); surveys that probe the cluster population at higher redshift are sensitive to the growth of density fluctuations, allowing one to break the ΩM -σ8 degeneracy that arises from local cluster abundance constraints (Eke et al. 1996; Bahcall & Fan 1998). Wang & Steinhardt (1998) argued that a measurement of the changes of cluster abundance with redshift would provide constraints on the dark energy equation of state parameter w ≡ p/ρ. Describing the problem in terms of cluster abundance only makes sense in the local universe , because, of course, one cannot measure the cluster abundance without know- ing the survey volume; the survey volume beyond z ∼ 0.1 is sensitive to cosmological parameters that affect the ex- pansion history of the universe -- namely, the matter den- sity ΩM , the dark energy density ΩE and the dark energy equation of state w. A cluster survey of a particular piece of the sky with appropriate followup actually delivers a list of clusters with mass estimates and redshifts -- that is, the redshift distribution of galaxy clusters above some de- tection limit. Recently, it has been recognized that with current in- strumentation it is possible to use such surveys of galaxy clusters extending to redshifts z > 1 to precisely study the amount and nature of the dark energy (Haiman et al. 2001). Clusters are promising tools for precision cosmolog- ical measurements, because they exhibit striking regularity and they exist throughout the epoch of dark energy dom- ination. Moreover, their use is complementary to studies of cosmic microwave background (CMB) anisotropy and SNe Ia distance measurements (Haiman et al. 2001; Levine et al. 2002; Hu & Kravstov 2002). Following Haiman et al. (2001), a series of analyses appeared that explore the theo- retical and observational obstacles to precise cosmological measurements with cluster surveys : Holder et al. (2001) applied the Fisher matrix formalism to the cluster survey problem and showed that high yield SZE cluster surveys can provide precise constraints on the geometry of the uni- verse through simultaneous measurements of ΩE and ΩM . Weller et al. (2001) demonstrated that future SZE surveys might constrain the variation of the dark energy equation of state w(z). Hu & Kravstov (2002) examined the effects of cosmic variance on cluster surveys as well as including the effects of imprecise knowledge of a more complete list of cosmological parameters. Levine et al. (2002) examined an X -- ray cluster survey, showing that a sufficiently large survey allows one to measure cosmological parameters and constrain the all -- important cluster mass -- observable rela- tion simultaneously. An important caveat to these works is that the authors assumed that the evolution of cluster structure with red- shift was perfectly known. In this paper, we examine the effects of uncertainties about cluster structural evolution on cosmological constraints from cluster surveys, finding that current survey projections that ignore this evolution uncertainty overstate the cosmological sensitivity of the survey. Furthermore, we examine the effects of survey fol- lowup to measure a cluster mass -- like quantity Mf , demon- strating that an appropriately designed survey can over- come this evolution uncertainty. In addition, our calculations underscore the importance of incorporating information from multiple observables 1Department of Astronomy, University Of Illinois, 1002 West Green St., Urbana, IL 61801 2Department of Physics, University Of Illinois, 1002 West Green St., Urbana, IL 61801 1 2 Majumdar & Mohr into future cluster surveys. Clusters of galaxies are dark matter dominated objects with baryon reservoirs in the form of an intracluster medum (ICM) and a galaxy popu- lation. Clusters can be found through the light the galaxies emit, the gravitational lensing distortions the cluster mass introduces into the morphologies of background galaxies, the X -- rays emitted by the energetic ICM, the distortion that the hot ICM introduces into the cosmic microwave background spectrum (SZE), and the effects that the ICM has on jet structures associated with active galaxies in the cluster. These methods are largely complementary, each having different strengths. It appears that X -- ray and SZE signatures of clusters are higher contrast observables than are weak lensing or galaxy light. That is, massive galaxy clusters are more prominent relative to the far more abun- dant lower mass halos and the large scale filaments when viewed with the SZE and X-ray; projection effects are a far more serious concern when using galaxy light or weak lens- ing signatures. Studies of the highest redshift galaxy clus- ters will likely be done with the SZE, because of the red- shift independence of the spectral distortion in the CMB. Any effort to carry out a precise cosmological study using galaxy clusters will undoubtedly be most effective through some combination of these complementary, cluster observ- ables. The paper is arranged in the following way. In §2 we describe two representative surveys and survey followup. Section 3.1 contains a description of our estimates of the survey sensitivity when followup is included as well as a description of our fiducial model. Results are presented in §4 and discussed further in §5. 2. FUTURE GALAXY CLUSTER SURVEYS A study of using cluster surveys to probe dark energy begins with the redshift distribution of detectable clusters within a survey solid angle ∆Ω, dN dz = ∆Ω dV dzdΩ dn(M, z) dM (1) (z)Z ∞ f (M ) dM 0 comoving volume the where dV /dzdΩ is element, (dn/dM )dM is the comoving density of clusters of mass M , and f (M ) is the cluster selection function for the sur- vey. In this analysis we take f (M ) to be a step function at some limiting mass Mlim, which corresponds to the mass of a cluster that lies at the survey detection threshold. We use the cluster mass function dn/dM determined from structure formation simulations (Jenkins et al. 2001). In practice surveys select clusters using observables like the X -- ray flux, SZE flux, galaxy light or weak lensing shear. Thus, in addition to the ingredients above, one requires a virial mass -- observable relation (like M -- Lx, M -- Lsz or M -- γt). Low redshift clusters do exhibit regularity (i.e. David et al. 1993; Mohr & Evrard 1997), suggesting that observables like the ICM X-ray luminosity and tem- perature are good mass estimators (Finoguenov et al. 2001; Reiprich & Bohringer 2002). Cluster mass to light ratios have been studied for decades, and it may be that this body of work together with modern datasets will allow more conclusive statements about how well galaxy light traces cluster halo mass. Hydrodynamical simulations lead us to expect that the SZE luminosity (related to the to- tal thermal energy within the virial region) should be the best ICM observable for predicting mass, but we await new observations with next generation SZE instruments to demonstrate this. A central feature of these mass -- observable relations is that they evolve with redshift due to the increasing density of the universe at earlier times (and the changing ratio of distance to lense and source in the case of weak lensing). Within standard structure formation models, galaxy clus- ters form self -- similarly, and so there are standard evolu- tion models for each mass -- observable relation (e.g. Bryan & Norman 1998; Mohr et al. 2000; Evrard et al. 2002). Results to date suggest that the degree of cluster regu- larity locally and at intermediate redshift is comparable (Mushotzky & Scharf 1997; Mohr et al. 2000; Vikhlinin et al. 2002). However, given the central importance of cluster mass estimates in using surveys to study dark en- ergy, we can only regard these standard structure forma- tion models as a guide; ultimately, one needs to determine the evolution of cluster structure observationally. In this paper we examine the effects that non -- standard redshift evolution of cluster structure would have on our ability to use cluster surveys to study the dark energy. As de- tailed below, we explore non -- standard redshift evolution by allowing an additional dependence of (1 + z)γ in the evolution of the relevant mass -- observable relations. 2.1. An X -- ray and an SZE Survey To examine these new effects, we adopt two representa- tive surveys that are being promoted as ways of measuring the dark energy equation of state. Namely, we examine the following two high yield surveys: (i) a 104 deg2 flux lim- ited X -- ray survey proposed as part of the DUET mission to the NASA Medium -- class Explorer Program, and (ii) a 4,000 deg2 SZE survey to be carried out with a proposed 8m South Pole Telescope (SPT). Figure 1 contains a plot of the redshift distribution and limiting mass for both sur- veys. We model the DUET X-ray survey as having a bolo- metric flux limit of fx > 1.25 × 10−13 erg/s/cm2 (corre- sponding to fx > 5 × 10−14 erg/s/cm2 in the 0.5:2 keV band). For our fiducial cosmological model (see §3.2 be- low) this survey yields ∼21,600 detected clusters, consis- tent with the known X-ray log N -- log S relation for clusters (e.g. Gioia et al. 2001). For our mass -- observable relation, we adopt a bolometric X-ray luminosity -- mass relation fx(z)4πd2 L = AxM βx 200E2(z) (1 + z)γx (2) where fx is the observed flux in units of erg s−1cm−2, dL is in units of Mpc, M200 in units of 1015M⊙ is the mass enclosed within a radius r200 having a overdensity of 200 with respect to critical and H(z) = H0E(z). where E2(z) = ΩM (1 + z)3 + Ωk(1 + z)2 + Ω3(1+w) . The E(z) factors follow the evolution of the critical density of the universe ρcrit = 3H 2/8πG. We convert M200 to M (z), the halo mass appropriate for our mass function at redshift z using a halo model (Navarro et al. 1997, hereafter, NFW; see discussion below regarding the effects of uncertainties in this conversion). Our standard evolution model ignores the T 1/2 dependence of the bolometric bremsstrahlung ra- diation, because X -- ray surveys detect clusters using de- tected photons rather than detected energy. We intro- duce the possibility of non-standard evolution of the mass -- observable relation with the parameter γx. We take γx = 0 E Cluster Evolution, Surveys and Dark Energy 3 to be consistent with the observed weak evolution in the luminosity -- temperature relation (Vikhlinin et al. 2002), and we choose βx = 1.807 and log Ax = −3.926, consis- tent with observations (Reiprich & Bohringer 2002). The overall h scaling of the limiting mass is h−1.11. We model the SPT SZE survey as a flux limited survey with fSZ > 5 mJy at 150 GHz. Within our fiducial cos- mological model this survey would yield ∼ 13, 500 clusters with measured fluxes. The mass -- observable relation is fsz(z, ν)d2 A = 3.781f (ν)fICM T M200 (1 + z)γsz M200 = Asz (kBT )βsz E(z) (3) where f (ν) is the frequency dependence of the SZE distor- tion, fsz is the observed flux in mJy, T is in Kelvin, M200 is in units of 1015M⊙, fICM = 0.12 (e.g Mohr et al. 1999) and dA is in units of Mpc (see Diego et al. 2002). We use log Asz = 13.466, βsz = 1.48 (Finoguenov et al. 2001) and γsz = 0 to model standard structure evolution. In this form, the overall h scaling of the limiting mass is h−1.61. Note that in determining the estimated uncertainties on cosmological parameters, we allow the normalization of these mass -- observable relations to be free to vary. The survey contains enough information to solve for the best normalization and the cosmological parameters simultane- ously; therefore, shifts in model inputs like fICM within the observational uncertainties have minimal effect on our conclusions. Fig. 1. -- The cluster redshift distribution (heavy line) and mass limit (light line) of the 104 deg2 DUET X -- ray survey (dashed) and the 4, 000 deg2 South Pole Telescope (SPT) SZE survey (solid). The surveys are flux limited (fx > 1.25 × 10−13erg/s/cm2 and fsz > 5mJy at 150 GHz), and we impose a minimum cluster mass of 1.54 × 1014 M⊙. A generic problem with flux limited surveys is that at low redshift the implied mass limit drops well below those masses corresponding to galaxy clusters. The flux from a nearby object is spread over a much larger portion of the sky, and surface brightness selection effects become important. We model these complications by imposing a minimum cluster mass of 1014h−1 M⊙. This lower limit on the survey mass limit is readily apparent below z ∼ 0.25 in Fig 1. 2.2. Followup of Large Solid Angle Surveys The redshift distribution of clusters contains far more cosmological information than the surface density of clus- ters (Haiman et al. 2001) or the angular correlation func- tion (e.g. Komatsu & Seljak 2002). Thus, in both these surveys each detected cluster will be followed up with multi -- band optical and near-IR photometry to provide photometric redshift estimates. These same data can be used to estimate cluster masses through their weak lensing effects on background galaxies (e.g. Bartelmann & Schnei- der 2001) and the total detected light from cluster galaxies. In addition, some of these clusters can be followed up with detailed X -- ray, SZE or galaxy spectroscopic obser- vations that allow one to measure the mass-like quantity Mf (θ) = M (θ)/dA, which we will refer to as the followup mass. As an example, in the case of followup X -- ray obser- vations that deliver the projected ICM temperature profile and surface brightness profile, it is straightforward to ex- tract the underlying ICM density ρ(dAθ) and temperature profile T (dAθ) to then estimate the followup mass Mf as Mf (θ) = −θ (4) kBT (θ) Gµmp (cid:18) d ln ρ d ln θ + d ln T d ln θ(cid:19) where mp is the proton mass, kB is Boltzmann con- stant, G is Newton constant, and the ICM number den- sity n ≡ ρ/µmp. Note that only the shape of the ICM density and temperature profiles is required (i.e. knowl- edge of the actual distance to the system is not required). The followup -- mass Mf can then be examined at within some angle θ along with the cluster X-ray or SZE flux. At fixed redshift, this followup would produce an Mf -- fx of Mf -- fsz relation which would provide direct constraints on the structural evolution of the clusters. As we will see, the parameter sensitivity of these scaling relation observa- tions can exhibit quite different degeneracies than for the cluster redshift distribution, making the scaling relations and the cluster redshift distribution complementary. In §3.1 below, we describe how these survey followup obser- vations are included in our estimates of the cosmological sensitivity of the survey. 3. COSMOLOGICAL SENSITIVITY OF A SURVEY 3.1. Fisher Matrix Technique We employ the Fisher matrix technique to probe the relative sensitivities of two cluster surveys to different cos- mological and cluster structural parameters. The Fisher matrix information for a data set (see Tegmark et al. 1997; Eisenstein et al. 1998) is defined as Fij ≡< ∂ 2 ln L >, ∂pi∂pj where L is the likelihood for an observable ( dN dz for the survey and Mf for the followup) and pi describes our pa- rameter set. The inverse F −1 ij describes the best attainable covariance matrix [Cij ] for measurement of the parameters considered. The diagonal terms in [Cij ] then give the un- certainties on each of our parameters. In calculating these uncertainties, we have added the Fisher matrix for the fol- lowup observations (F f ij,), the Fisher matrix for the cluster redshift distribution (F s ij ) and several external priors that will be discussed below. We construct the survey Fisher matrix F s ij following Holder et al. (2001) as F s ij = Σn ∂dN/dz ∂dN/dz 1 ∂pi ∂pj dN/dz , (5) 4 Majumdar & Mohr Table 1 Estimated Parameter Constraints Description Priors SPT SZE Survey Std Evolution Non-Std Evolution + 1% Followup + 10% Followup + 100% Followup + Flat + 100% F-up 100% F-up Only DUET X -- ray Survey Std Evolution Non-Std Evolution + 1% Followup + 10% Followup + 100% Followup + Flat + 100% F-up 100% F-up Only ΩM 0.0177 0.0184 0.0167 0.0147 0.0139 0.0139 0.2360 0.0147 0.0255 0.0157 0.0139 0.0134 0.0133 0.2018 Ωtot 0.0100 0.0077 0.0079 0.0074 0.0070 0.0068 - 0.0099 0.0087 0.0092 0.0086 0.0085 0.0084 - 0.0100 σ8 w h n ΩB log A β γ 0.0134 0.0195 0.0157 0.0122 0.0109 0.0109 - 0.0113 0.0441 0.0167 0.0116 0.0105 0.0100 - 0.1629 0.2487 0.1404 0.1237 0.1207 0.0406 0.6369 0.1659 0.4505 0.1655 0.1454 0.1414 0.0593 0.5500 0.0323 0.0500 0.0040 0.0323 0.0323 0.0285 0.0259 0.0248 0.0235 - 0.0323 0.0324 0.0233 0.0201 0.0191 0.0187 - 0.0478 0.0479 0.0476 0.0472 0.0471 0.0439 - 0.0476 0.0476 0.0475 0.0475 0.0473 0.0467 - 0.0040 0.0040 0.0040 0.0040 0.0040 0.0039 - 0.0040 0.0040 0.0040 0.0040 0.0040 0.0039 - 0.1392 0.1505 0.1423 0.1320 0.1004 0.1001 0.2204 0.0928 0.2216 0.1060 0.0814 0.0586 0.0585 0.1408 0.0064 0.0064 0.0064 0.0061 0.0046 0.0046 0.0067 0.0060 0.0096 0.0062 0.0055 0.0040 0.0040 0.0056 - 0.4602 0.2248 0.0958 0.0353 0.0352 0.1912 - 1.2955 0.3310 0.1185 0.0444 0.0444 0.1899 where we sum over n redshift bins of size ∆z = 0.01 to zmax = 3.0 and dN/dz represents the number of surveyed clusters in each redshift bin. The Fisher matrix for the followup is constructed as F f ij = Σn dV dz Z dM f dn dM ∂Mf ∂pi ∂Mf ∂pj 1 σ2 Mf! (6) where Mf is a function of halo mass M and angular ra- dius θ, and f (dn/dM ) represents the number of clusters of mass M for which followup mass measurements are avail- able in a particular redshift bin. We examine cases where the followup fraction f is 1%, 10% and 100%. To generate the followup Fisher matrix, we calculate the cluster binding mass within radius r = dAθ for a halo with virial mass M . To do this calculation we assume cluster mass profiles are well represented, on average, by NFW models with concentration index c = 5. In practice clusters undergo merging quite frequently and there is a range of halo shapes. This introduces a "theoretica"l un- certainty to the followup mass. In this analysis we take σMf = 0.3Mf to be the characteristic uncertainty in the followup mass measurments. This uncertainty reflects the observational uncertainty on individual cluster followup mass measurements as well as the uncertainties inherent in predicting the followup mass from the halo virial mass. As is clear from Eqn 6, the redshift and mass distribu- tions of the followup clusters match those of the full cluster survey sample; that is, we don't choose to followup only high redshift clusters, which would presumably provide the tightest constraints on our evolution parameter. We also choose θ to be a dynamically varying quantity fixed to be 95% of the virial radius corresponding to the cluster lim- iting mass at each redshift. Thus, followup at all redshifts corresponds to mass -- like measurements at radii within the virial radius (rθ(M ) < r200(M )). 2001), ΩB = 0.047 (Burles & Tytler 1998), and a COBE normalized σ8 = 0.72 (Bunn & White 1997). Note that we use a rather low value of σ8, which is consistent with the recent 2dF analysis (Lahav et al. 2002). Because the expected number density of clusters is very sensitive to the value of σ8, our fiducial SZE survey has fewer clusters when compared to some previous studies (i.e. Holder et al. 2001). Cosmological constraints from cluster surveys are com- plementary to constraints from SNe Ia distance measure- ments and observations of the anisotropy of cosmic mi- crowave background. This is particularly true when it comes to using cluster surveys to measure the dark en- ergy equation of state parameter w (Haiman et al. 2001). In combination with precise CMB constraints on the cur- vature (Ωk = 0), cluster surveys enable precise measure- ments of the dark energy equation of state; however, when curvature is allowed to depart from zero -- even slightly -- the cluster constraints on w weaken considerably. For a flat universe , the constraints from our SZE survey assum- ing standard evolution and a 100% followup on w (ΩM ) are 0.0406 (0.0139), whereas for σk = 0.01 the constraints are 0.1207 (0.0139). For the analysis presented here, we adopt relatively con- servative priors from future CMB anisotropy studies and distance measurements. We assume the power spectrum index n will be known to 5% , i.e. σn = 0.05, the Hubble parameter will also be known to 5%, i.e σh = 0.0325 and the total density parameter Ωtot will be known to 1%, i.e σk = 0.01. In addition, we take the prior on the baryon density parameter (ΩB = 0.047) to be σΩB = 0.004. For reasonable values of ΩB, surveys are affected by ΩB vari- ations only through minor effects on the transfer function for density perturbations (see also Levine et al. 2002). Fi- nally, we neglect the possibility of a variation with redshift in the equation of state parameter w (Weller et al. 2001). 3.2. Fiducial Cosmology and External Constraints 4. COSMOLOGICAL PARAMETER CONSTRAINTS The fiducial cosmological parameters of our model are h = 0.65 (i.e. Hendry et al. 2001; Ajhar et al. 2001; Reese et al. 2002), ΩM = 0.3 (i.e. Mohr et al. 1999; Grego et al. 2001), Ωtot = ΩM + ΩE = 1 (i.e. Netterfield et al. 2001; Pryke et al. 2001), w = −1, n = 0.96 (Netterfield et al. Our results are listed in Table 1 and highlighted in the following figures. Table 1 contains a listing of 1σ uncertainties on all seven cosmological and three mass- observable relation parameters (see Eqns 2 & 3 for def- initions). The first line contains a listing of the priors Cluster Evolution, Surveys and Dark Energy 5 adopted for the runs. Following that are the results for the SZE survey and then the X-ray survey. For each sur- vey we show the constraints in the case of standard evo- lution (i.e. γ = 0) followed by non -- standard evolution (γ is free parameter). The lines that follow highlight the impact of 1%, 10% and 100% followup. Following those scenarios is what we consider to be the ideal case of a flat universe with 100% survey followup. Finally, we show the case for followup only (i.e. the Fisher matrix derived from the redshift distribution isn't used). The constraints in this line provide some insights into the parameter lever- age that is afforded by cluster followup observations. In all cases the uncertainties are absolute (i.e. σM = 0.0177 means ΩM = 0.3 ± 0.0177). Fig. 2. -- Constraints on w and ΩM for an SZE (above) and an X-ray survey (below). The star marks the fiducial model. Con- tours denote joint 1σ constraints in five scenarios: constraints from dN/dz where (i -- long dashed) the cluster evolution is unknown; con- straints from dN/dz and followup mass measurements for (ii -- short dashed) 1% of sample, (iii -- dotted) 10% of sample, and (iv -- solid) 100% of sample. The case for 100% followup plus added prior of a flat universe is also shown (v -- dot -- dashed). The followup mass mea- surements are estimated to have fractional uncertainties of 30%. 4.1. Importance of Non-Standard Evolution In the standard evolution case, the SZE and X -- ray sur- veys compare favorably, yielding 1σ absolute uncertain- ties on w (ΩM ) of 0.1629 and 0.1659 (0.0177 and 0.0147), respectively. However, when one takes into account the possibility of non-standard evolution, the constraints on w weaken by almost a factor of 2 to ∼ 0.25 for SZE and a factor 3 to ∼ 0.45 for X -- ray; ΩM constraints weaken by close to a factor of 2 to ∼ 0.026 for the X -- ray but are only slightly affected in the SZE survey. The constraints from the cluster redshift distribution dN/dz on γsz/x are very weak at 0.46 and 1.29, respectively; this large un- certainty in the evolution of the mass -- observable relation contributes to the weakened sensitivity to other cosmolog- ical parameters. The importance of evolution in interpreting the cluster redshift distribution contrasts somewhat with the results of the Levine et al. (2002) study, which showed that prior knowledge of the normalization of the mass -- observable re- lation has only a weak effect on the cosmological sensi- tivity of cluster surveys (see also Diego et al. 2001). In their study, they only considered the standard evolution model. Within the context of uncertain evolution of the mass -- observable relation one needs observations in addi- tion to dN/dz to determine the evolution parameter γ and regain high sensitivity to the equation of state parameter w. Next we examine the effects of including followup mass measurements. 4.2. Effects of Followup Mass Measurements Figure 2 contains joint constraints on Ωm and w for the two surveys that highlight the effect of survey followup. For each survey we show constraints with non-standard evolution and no followup (long dashed), 1% (dashed), 10% (dotted), and 100% (solid) followup along with 100% followup in a flat universe (dot -- dashed). The figure makes clear that even a limited followup program can greatly im- prove cosmological constraints. Table 1 shows that as the followup fraction increases, the constraints on w in the SZE survey tighten from 0.25 (no followup) to 0.14 (1% followup) to 0.12 (10% and 100% fol- lowup). The difference between 10% and 100% followup is minimal, suggesting that the ten times higher cost of full survey followup is not a worthwhile investment when viewed solely from the perspective of obtaining constraints on the equation of state of the dark energy. Notice that the constraints on w in the cases of even limited followup are somewhat better than the constraints in the cases where we assume complete knowledge of the evolution of the mass -- observable relation. In the X-ray survey, a program to followup as few as 1% of the clus- ters can offset the increase in uncertainties that we see in going from standard evolution to non-standard evolution. In the SZE survey, 1% followup produces constraints that are somewhat better, reducing the uncertainty on w from 0.1629 in the standard evolution scenario to 0.1404 in the non-standard evolution + 1%followup. This can be traced to our assumption in these calculations that the redshift distribution of followup mass measurements matches the redshift distribution of the full survey. The higher redshift followup measurements contain more information about evolution, and the SZE survey probes to higher redshift than does the X -- ray survey (see Fig. 1) Table 1 contains a listing of the effects of followup on all parameters. It is clear that followup mass measure- ments dramatically reduce the projected uncertainties on cosmological and scaling relation parameters. As is ev- ident from the last column in the table, even a modest followup of 10% of the clusters reduces the uncertainty on γ from 0.46 to 0.10 for the X-Ray case and 1.29 to 0.12 for the SZE survey. With full followup, the constraint on 6 Majumdar & Mohr the mass -- observable relations, more followup is better. w shrinks from 0.25 to 0.12 in the SZE and 0.45 to 0.14 in the X -- ray survey. Even with followup of only 1% of the clusters in the SZE survey, one reduces the uncertainty on w by half. Comparison of the uncertainties on γ for non- standard evolution with no followup to the followup only case shows clearly that followup is very effective in con- straining γ. For example, for the X -- Ray case the followup itself can constrain γ to ∼ 0.19 compared to 1.29 that one can achieve from the survey only. This underscores the ad- vantage of having a followup program for cluster surveys. Fig. 4. -- The sensitivity of the Followup (above) and Survey (be- low) to change in w is shown for both X -- Ray and SZE surveys. The four cases are: (i -- solid) SZE survey weighted sensitivity of Fij to change in w and (ii -- dashed) X -- Ray survey weighted sensitivity Fij to change in w (see Eqn 5 & Eqn 6 for an explanation of Fij); (iii -- dot-dashed) same as (i) but per unit cluster and (iv -- dotted) same as (ii) but per unit cluster. The redshift distribution of the clusters are shown in. 4.3. Redshift Variation of Parameter Sensitivity In Fig 4 we display an estimate of the redshift variation of the survey and followup sensitivity to w. We do this by examining the derivative with respect to redshift of the w-w components of the survey and followup Fisher matri- ces. These are shown using heavy lines for both the X-ray (dashed) and SZE (solid) surveys. We also show an esti- mate of the per unit cluster sensitivity using lighter lines for the X-ray (dot -- dashed) and SZE (dotted) surveys. The axes for the heavy curves are included on the left of the figure, and the axes corresponding to the per unit clus- ter values are located on the right of the figure. In both the X-ray and SZE cases and for both followup and the survey itself, the heavy curves show sensitivity that peaks at lower redshift than in the lighter curves that show the per unit cluster sensitivity. This is simply a reflection of the redshift distributions of the clusters detected in the surveys, which peak at z < 0.5 in both surveys. Consider now the lower panel. The higher redshift na- ture of the SPT SZE survey relative to the DUET X-ray survey is clear in this panel, where we see that the w sen- sitivity of the SZE survey peaks near z ∼ 1 (solid line). For the X-Ray survey, the w sensitivity peaks at z ∼ 0.7 (dashed line). However, it is clear that the sensitivity per unit cluster ((dF S ww/dz)/(dN/dz), denoted by dotted line Fig. 3. -- Constraints on the mass -- observable relation normal- ization A and redshift evolution (1 + z)γ for an SZE (above) and an X-ray survey (below). The star marks the fiducial model. Con- tours denote joint 1σ constraints in four scenarios: constraints from dN/dz where (i -- long dashed) the non standard cluster evolution is unknown; constraints from dN/dz and followup mass measurements for (ii -- short dashed) 1% of sample, (iii -- dotted) 10% of sample, and (iv -- solid) 100% of sample. The followup mass measurements are estimated to have fractional uncertainties of 30%. In Fig. 3 we show the constraints on the mass-observable relation normalization A and the evolution parameter γ for the four cases: no followup (long -- dashed) and a followup of 1% (short -- dashed), 10% (dotted) and 100% (solid) of the clusters. Followup has strikingly different effects in the SZE and X-ray surveys. Followup in the SZE survey is more effective at constraining the evolution parameter γsz, due to the greater redshift depth of this survey. The differences in the constraints on log10 A generally reflect the different definitions of the normalization and its rela- tionship to halo mass (see Eqns 2 & 3). In contrast to the previous figure that shows the effects of followup on w constraints, it is clear from this figure that if one really wants to understand the normalization and evolution of Cluster Evolution, Surveys and Dark Energy 7 for SZE and dot-dashed line for X-Ray) increases as we go to higher redshift. This emphasizes the importance of high redshift cluster surveys for probing w. High redshift clusters in the DUET survey are even more sensitive to w, because they are more massive and lie well beyond the exponential cutoff in the mass function; however, these clusters are so rare that none are detected in the DUET survey. The sharp cutoff and first mini -- peak at redshift z ∼ 0.3 of dF S ww/dz for the survey is a direct result of our re- quirement that clusters have masses above 1014h−1M⊙. In Fig 1 the plot of the limiting mass becomes flat at red- shifts below z ∼ 0.3; we include this mass cutoff for several reasons: (1) the mass -- observable scaling relations we've adopted are for clusters, and they are inappropriate for group scale systems and (2) a flux limited survey becomes sensitive to surface brightness limitations at low redshift, where the flux from these nearby objects is spread over a larger and larger solid angle. Introducing a minimum mass in our survey causes this interesting artifact in the w sen- sitivity at low redshift, which can be understood by con- sidering the competitive behaviour of the w-dependences of the limiting mass, the survey volume element, and the growth factor of density perturbations. As long as the limiting mass is constant, the opposite sensitivities of the volume element and the growth factor to w determine the net w-sensitivity of the survey. At low redshift, the w- sensitivity of the volume element dominates over that of the growth factor. Beyond redshift ∼ 0.3, the limiting mass suddenly rises above the minimum, allowing the net w sensitivity to include that of the limiting mass. The lim- iting mass is sensitive to w primarily through the angular diameter or luminosity distance to the redshift of inter- est, and this sensitivity combines with that of the growth factor to offset to a larger degree the w sensitivity of the volume element. Note that dF S ww/dz is positive definite which is the reason for the visual appearance of a break in the sensitivity, which is actually reflecting an underlying change in sign of the w sensitivity. For the followup (upper panel), the sensitivity per unit cluster peaks at z ∼ 0.8 for both the surveys. However, survey weighted sensitivities of the followup to variation in w are peaked at much lower redshifts, where the sur- vey yields are much higher. This is due to the fact that the w-sensitivity of the followup is a balance between the w-sensitivities of the mass observable and the number of clusters one has at that redshift to make the measure- ment. The fact that the redshift dependence of the cluster mass observable is similar for both X -- ray and SZE just reflects the weak mass dependence of this sensitivity. Sur- vey strategists should consider a followup program that targets predominantly redshift z ∼ 1 clusters if their goal is to constrain the evolution parameter and improve con- straints on w; however, having evolution information over the entire redshift range of the survey is critical to testing the form of the non-standard evolution model, which we parametrize here simply as (1 + z)γ. 5. DISCUSSION AND CONCLUSIONS Any attempt to precisely measure the dark energy equa- tion of state w with cluster surveys will require (i) a strong external prior on the curvature (presumably from CMB anisotropy studies) and (ii) an understanding of the evolu- tion of the relation between cluster halo mass and observ- able properties like the X -- ray luminosity, SZE luminos- ity, galaxy light or weak lensing shear. We have examined the effects of current uncertainties about cluster structural evolution; for two recently proposed cluster surveys the estimated constraints on w are ∼2 -- 3 times weaker than if one assumes full knowledge of cluster evolution. Con- straints on other interesting cosmological parameters are also weakened (see Table 1). Followup observations to measure cluster masses di- rectly will enable one to solve for cluster structure evo- lution and to enhance cosmological constraints. We have examined the effects of followup mass observations from hydrostatic or dynamical methods, and we find that even modest followup of 1% of the cluster sample can improve survey constraints. Full followup with mass measurements that are 30% uncertain, on average, provide cosmological constraints that match or surpass those possible through dN/dz alone with full knowledge of cluster evolution. Full followup with weak lensing mass measurements is cur- rently being planned for the SPT SZE survey. The implications are quite interesting. Essentially, to do precision cosmology with cluster surveys and followup, we need only know that clusters conform to mass -- observable scaling relations, and that these relations evolve in some well behaved manner. Then, together with our well estab- lished theoretical framework for structure formation, the observed cluster redshift distribution and followup masses of as few as 1% of the sample then provide enough in- formation to deliver precise constraints on cosmological parameters and the character and evolution of the mass -- observable relation simultaneously. In a sense cluster sur- veys with limited followup are self -- calibrating: one gains detailed knowledge of the structure of the tracers (i.e. galaxy clusters) and detailed knowledge of the evolution of the universe from the same dataset. We have focused here on the mean equation of state pa- rameter w, and for the two surveys considered, we examine the redshift variation of the sensitivity to w. In the SPT SZE survey, the sensitivity to the dark energy equation of state peaks at z ∼ 1 with sensitivity at or above half the peak for 0.65 ≤ z ≤ 1.5. In the DUET X-ray survey, the sensitivity peaks at z ∼ 0.7 with sensitivity at or above half the peak for 0.45 ≤ z ≤ 1.0. In the case of both surveys the w sensitivity of followup mass measurements peaks near redshift z ∼ 0.35. In the case of the clus- ter redshift distribution, the most information about w is provided by the highest redshift clusters, and so deeper, more sensitive surveys will in general be better for studies of the dark energy equation of state. One interesting feature of our analysis is the orientation of the elliptical constraints on w and ΩM (see Fig. 2). In general, the rotation of the parameter degeneracy can be understood as the result of competing effects of changes in the volume element and the growth factor as parameters vary. Variations in w (and ΩM ) affect the survey yield in different ways at different redshifts, and so the w-ΩM degeneracy depends on the redshift distribution of a par- ticular survey. Rotations of parameter degeneracies occur as the maximum redshift of the survey is varied (Levine et al. 2002; Hu & Kravstov 2002). We have further found that changing the prior on Ωtot and changing the degree 8 Majumdar & Mohr of mass followup on a survey also result in rotations of the parameter degenaracy. This behavior has interesting implications for the design of cluster surveys that are op- timally complementary to CMB anisotropy and SNe Ia distance measurements, and it deserves further study. In addition, we emphasize that the final constraint on the determination of cosmological parameters depends sensitively on the survey strategy and also the details of the followup. For example, a different definition of Mf (θ) would lead to slightly different uncertainties. Changing θ from a quantity that varies with redshift to some fixed value leads to modest variations of the constraints. In gen- eral, best results can be obtained by optimizing Mf (θ) so that one probes as mush of the virial mass as possible. SM thanks Ben Wandelt and Shiv Sethi for helpful conversations. JM thanks Zoltan Haiman for many fun discussions of cluster surveys. This work has been sup- ported by NASA Long Term Space Astrophysics grant NAG5 -- 11415 and Chandra X -- ray Observatory archival grant AR1 -- 2002X, awarded through the Smithsonian As- trophysical Observatory. REFERENCES Ajhar, E. A., Tonry, J. L., Blakeslee, J. P., Riess, A. G., & Schmidt, B. P. 2001, ApJ, 559, 584 Bahcall, N. A. & Fan, X. 1998, ApJ, 504, 1+ Bartelmann, M. & Schneider, P. 2001, Phys. Rep., 340, 291 Bryan, G. L. & Norman, M. L. 1998, ApJ, 495, 80 Bunn, E. F. & White, M. 1997, ApJ, 480, 6 Burles, S. & Tytler, D. 1998, ApJ, 507, 732 David, L. P., Slyz, A., Jones, C., Forman, W., Vrtilek, S. D., & Arnaud, K. A. 1993, ApJ, 412, 479 Diego, J. M., Mart´inez-Gonz´alez, E., Sanz, J. L., Benitez, N., & Silk, J. 2002, MNRAS, 331, 556 Diego, J. M., Mart´inez-Gonz´alez, E., Sanz, J. L., Cay´on, L., & Silk, J. 2001, MNRAS, 325, 1533 Eisenstein, D. J., Hu, W., & Tegmark, M. 1998, ApJ, 504, L57 Eke, V. R., Cole, S., & Frenk, C. S. 1996, MNRAS, 282, 263 Evrard et al. 2002, ApJ, in press (astro -- ph/0110246) Finoguenov, A., Reiprich, T. H., & Bohringer, H. 2001, A&A, 368, 749 Gioia, I. M., Henry, J. P., Mullis, C. R., Voges, W., Briel, U. G., Bohringer, H., & Huchra, J. P. 2001, ApJ, 553, L105 Grego, L., Carlstrom, J. E., Reese, E. D., Holder, G. P., Holzapfel, W. L., Joy, M. K., Mohr, J. J., & Patel, S. 2001, ApJ, 552, 2 Haiman, Z. ., Mohr, J. J., & Holder, G. P. 2001, ApJ, 553, 545 Hendry, M. A., Rauzy, S., Goodwin, S. P., & Gribbin, J. 2001, MNRAS, 324, 717 Henry, J. P. 1997, ApJ, 489, L1 Holder, G., Haiman, Z. ., & Mohr, J. J. 2001, ApJ, 560, L111 Hu, W. & Kravstov, A. 2002, ApJ, submitted astro-ph/0203169 Jenkins, A., Frenk, C. S., White, S. D. M., Colberg, J. M., Cole, S., Evrard, A. E., Couchman, H. M. P., & Yoshida, N. 2001, MNRAS, 321, 372 Komatsu, E. & Seljak, U. 2002, MNRAS, submitted (astro*ph/0205468) Lahav et al. 2002, MNRAS, in press (astro*ph/0112162) Levine, E. S., Schulz, A. E., & White, M. 2002, ApJ, submitted (astro*ph/0204273) Mohr, J. J. & Evrard, A. E. 1997, ApJ, 491, 38 Mohr, J. J., Mathiesen, B., & Evrard, A. E. 1999, ApJ, 517, 627 Mohr, J. J., Reese, E. D., Ellingson, E., Lewis, A. D., & Evrard, A. E. 2000, ApJ, 544, 109 Mushotzky, R. F. & Scharf, C. A. 1997, ApJ, 482, L13 Navarro, J. F., Frenk, C. S., & White, S. D. M. 1997, ApJ, 490, 493 Netterfield et al. 2001, ApJ, submitted (astro*ph/0104460) Pryke, C., Halverson, N. W., Leitch, E. M., Kovac, J., Carlstrom, J. E., Holzapfel, W. L., & Dragovan, M. 2001, ApJ, submitted (astro*ph/0104490) Reese, E. D., Carlstrom, J. E., Joy, M., Mohr, J. J., Grego, L., Holder, G. P., & Holzapfel, W. L. 2002, ApJ, in press Reiprich, T. H. & Bohringer, H. 2002, ApJ, 567, 716 Tegmark, M., Taylor, A. N., & Heavens, A. F. 1997, ApJ, 480, 22+ Viana, P. T. P. & Liddle, A. R. 1999, MNRAS, 303, 535 Vikhlinin, A., VanSpeybroeck, L., Markevitch, M., Forman, W. R., & Grego, L. 2002, ApJ, submitted (astroph/0207445) Wang, L. & Steinhardt, P. J. 1998, ApJ, 508, 483 Weller, J., Battye, R., & Kniessl, R. 2001, astro-ph (0110353)
astro-ph/0501484
1
0501
2005-01-22T02:46:53
A theoretical look at the direct detection of giant planets outside the Solar System
[ "astro-ph" ]
Astronomy is at times a science of unexpected discovery. When it is, and if we are lucky, new intellectual territories emerge to challenge our views of the cosmos. The recent indirect detections using high-precision Doppler spectroscopy of now more than one hundred giant planets orbiting more than one hundred nearby stars is an example of such rare serendipity. What has been learned has shaken our preconceptions, for none of the planetary systems discovered to date is like our own. However, the key to unlocking a planet's chemical, structural, and evolutionary secrets is the direct detection of the planet's light. I review the embryonic theory of the spectra, atmospheres, and light curves of irradiated giant planets and put this theory into the context of the many proposed astronomical campaigns to image them.
astro-ph
astro-ph
A theoretical look at the direct detection of giant planets outside the Solar System Adam Burrows Department of Astronomy, The University of Arizona, Tucson, AZ 85721 e-mail: [email protected] I. HEADING Astronomy is at times a science of unexpected discovery. When it is, and if we are lucky, new intellectual territories emerge to challenge our views of the cos- mos. The recent indirect detections using high-precision Doppler spectroscopy of now more than one hundred giant planets orbiting more than one hundred nearby stars is an example of such rare serendipity. What has been learned has shaken our preconceptions, for none of the planetary systems discovered to date is like our own. However, the key to unlocking a planet’s chemical, structural, and evolutionary secrets is the direct detection of the planet’s light. I review the embryonic theory of the spectra, atmospheres, and light curves of irradi- ated giant planets and put this theory into the context of the many proposed astronomical campaigns to image them. II. INTRODUCTION: THE NEWLY-DISCOVERED WORLDS Direct detection of an extrasolar planet requires that its dim light be separated from under the glare of its bright parent star. However, such high-contrast imaging (e.g., a part in 107−10 in the visible) has not to date been achieved. Instead, the vast majority of known extrasolar giant planets (EGPs) have been discovered from the ground using the indirect technique of high-precision stellar spectroscopy [1–3]. Due to gravitational attraction, an 1 orbiting planet induces a Doppler wobble in its parent star. If the planet is massive and close enough, the periodic variation in the stellar spectral lines can be measured. The planet’s period (P ), eccentricity (e), orbital semi-major axis (a), and projected mass (Mp sin(i)), where i is the inclination of the orbit, can thereby be determined. The larger Mp sin(i), the larger the signal. This is the reason the first planets detected were the EGPs. Terrestrial planets, such as Earth and Venus, are ∼300 times lighter than Jupiter, while ice giants, such as Uranus and Neptune, are ∼20 times lighter. Before I delve into the physical theory of EGPs and their direct detection, I summarize the basic facts of the known members of the EGP family. The first extrasolar giant planet culled was 51 Peg b [1] and it is in a tight 4.2-day orbit, one hundred times closer to its primary than is Jupiter to the Sun. To date, more than 140 EGPs/planets have been discovered, more than 25 of which are in more than 10 multiple systems. 55 Cancri houses a quadruplet [4], one of which has a mass near that of Neptune (∼17 Earth masses), υ And house a triplet, and GJ 876 houses a doublet in a two-to-one orbital resonance. (We follow the convention by which the planet’s name is given by the star’s name, with an appended lower-case letter, either b, c, or d, in discovery order.) The projected masses of the known Doppler planets vary from ∼0.06 (!) MJ to above 10 MJ , where MJ is a Jupiter mass, which is 318 Earth masses or roughly 10−3 solar masses. The more massive objects may be brown dwarfs with a different provenance (see Box). Radial-velocity (Doppler) techniques can not distinguish EGPs and Neptune-mass planets from brown dwarfs. The orbital periods of the known EGPs span a vast range from ∼1.2 days to ∼12 years, their semi-major axes extend from ∼0.022 AU to ∼6.0 AU, where an AU is an Astronomical Unit, the distance between the Earth and the Sun, and their orbital eccentricities vary from 0.0 to above 0.9. For comparison, Jupiter resides 5.2 AU from our Sun, has an orbital period of ∼12 years, and has an orbital eccentricity of ∼0.05. Table 1 provides these basic data for a representative subset of the current EGP bestiary. The extremely close-in EGPs, such as 51 Peg b, τ Boo b, HD209458b, and OGLE-TR56b [5–7], were a surprise, but no less so than was the heterogeneity of the masses and orbital 2 properties of the emerging EGP family. To be sure, the Doppler technique selects for the closer representatives, but they must exist to be detected. As would be expected due to tidal dissipation, the close-in EGPs with orbital distances smaller than ∼0.06 AU all have nearly circular orbits. There seems to be a correlation between the probability of finding an EGP and the metallicity of its parent star. The “metallicity” of a star is the mass fraction of elements, such as carbon, oxygen, nitrogen, neon, magnesium, silicon and iron, that are heavier than helium. Hydrogen and helium predominate in stars and giant planets, comprising ∼98% by mass of the Sun. The more super-solar the heavy-element composition of the potential parent, the more likely we are to find an EGP in orbit. This may be a hint concerning the processes of giant planet formation, and is in keeping with the 3-5×solar excesses measured in Jupiter and Saturn. The current census reveals that there is a ∼5% a priori chance of finding a giant planet by the Doppler technique around a nearby (< ∼ 50 parsecs ≡ 160 light- years) star, but a ∼20% chance of finding one around a star with at least twice the Sun’s metallicity (J. Valenti & D.A. Fischer, in preparation). Presumably, the inclinations of EGP orbits are distributed randomly on the sky. Hence, the probability that the orbit is edge-on (i = 90◦) is approximately R∗/(2a), where R∗ is the stellar radius. Given this, the close-in EGPs have the largest chance of transiting the stellar disk, during which time the star will dim by a fraction (Rp/R∗)2, where Rp is the planet’s radius. Since RJ (the radius of Jupiter, ∼7.14 × 104 kilometers) is roughly 10% of the radius of the Sun, this ratio is expected to be roughly 1%. A 1% dimming is easily detectable from the ground. At a=0.045 AU and a distance (d) of 47 parsecs, the planet around the F8V/G0V star HD209458 was the first of only a handful of EGPs that are now known to transit their primaries and a periodic dimming at the ∼1.6% level was measured [8–10]. The transit of HD209458 lasts ∼3 hours (out of a total period of 3.524738 days). This was followed by the photometrically-selected transiting EGPs OGLE-TR-56b, OGLE-TR- 113b, OGLE-TR-132b, OGLE-TR-111b, and TrES-1 [5–7,11–13]. Many more EGP transits are anticipated during the Kepler [14] and Corot [15] space missions. These projects are 3 focussed on detecting transits around a fraction of the tens of thousands of stars they will monitor and will boast photometric accuracy (∼10−5) sufficient to measure not only transits by EGPs, but by Earth-like planets. The import of an EGP transit lies in the simultaneous measurement of both the orbital inclination (and, hence, with Doppler spectroscopy, the mass) and the radius of the planet. Knowledge of Rp and Mp (with some knowledge of the star) can be used to constrain theories of the structure and evolution of the close-in EGP [16–18]. Currently, non-transiting EGPs are mute concerning such physical information. HD209458 is close and bright enough that the STIS instrument on HST was used not only to obtain photometric precision of ∼0.01% [10], but to distinguish a difference at the 4-σ level in the planetary transit radius in and out of the Na-D line at 0.589 µm . In this way, neutral sodium atoms were discovered in HD209458b’s atmosphere [19–21]. Though indirect, this is the first measurement of the composition of the atmosphere of an extrasolar planet. Since then, the Lyman-α line of hydrogen has similarly been detected in HD209458b’s atmosphere [22], and by the large magnitude (∼15%) of the photometric dip at this UV wavelength (λ) a planetary wind [23] comprised of molecular break-up products has been inferred. However interesting, transits are rare and no substitute for direct imaging and optical and infrared spectra. Spectra can provide diagnostics for atmospheric composition, radius, gravity, and mass. Images are ground truth for the existence of a planet and provide orbital information that complements that gleaned from Doppler measurements. Furthermore, direct detection might be able to distinguish the different models of giant planet formation, such as nucleation around an ice/rock core [24] and direct collapse [25], and can probe the outer orbits where the majority of EGPs might reside. Since the indirect radial-velocity technique for EGP discovery selects for the closer vari- ety, it is likely that a large reservoir of giants exists at distances and orbital periods beyond the reach of Doppler spectroscopy. Furthermore, the best theory for the orbits of the closest EGPs is that they migrated in from further out during the early phase of star and planet for- mation [26]. This too would imply that a large pool of EGPs resides at larger separations. Indeed, it may be that the majority of stars in the solar neighborhood harbor planetary 4 systems, that only new techniques can reveal. This is where the direct planet detection methods, most effective at large angular distances from the parent star, will come into their own. III. THEORETICAL ATMOSPHERES AND CHEMISTRY OF EGPS After formation, without any significant internal sources of energy, an EGP gradually cools and shrinks. Its rate of cooling can be moderated by stellar irradiation, or by hydro- gen/helium phase separation when old and light [27], but is inexorable. Jupiter itself is still cooling and its total infrared plus optical luminosity is about twice the power intercepted from the Sun. The rate of cooling is a function of mass and composition, with more massive EGPs cooling more slowly. Hence, the instantaneous state of an EGP is a function of mass, age, composition, orbital distance, and stellar type, not just mass and composition. Unlike a star, EGP atmospheric temperatures are sufficiently low that chemistry is des- tiny. This is a distinguishing characteristic of substellar-mass objects (SMOs). The at- mosphere of a gaseous giant planet is the thin outer skin of molecules that regulates its emission spectrum and cooling rate. Molecular hydrogen (H2) is the overwhelming con- stituent, followed by atomic helium. An EGP’s effective temperature (Teff , the temperature of its “photosphere”) can vary from ∼1500 K for the more massive EGPs at birth to ∼50 K for the least massive EGPs after a Hubble time. This wide range translates into a rich variety of atmospheric constituents that for a given mass and elemental composition evolves significantly. At birth, Jupiter had a Teff near 600-1000 K and the appearance of a T dwarf [28] brown dwarf. It had no ammonia or water clouds and, due to the presence of atomic sodium in its hot atmosphere, had a magenta color in the optical [29]. Its atmosphere was depleted of aluminum, silicon, iron, calcium, and magnesium due to the formation and settling to depth of the refractory silicates (“dirt”) that condense in the temperaure range ∼1700-2500 K [30–32]. Water vapor (steam) was the major reservoir of oxygen, gaseous methane was the major reservoir of carbon, gaseous ammonia and molecular nitrogen were 5 the reservoirs of nitrogen, and H2S was the reservoir of sulfur. As it cooled, the layer of alkali metals was buried below the photosphere to higher pressures, but gaseous H2, H2O, NH3, and CH4 persisted to dominate the atmospheric composition. At a Teff of ∼400 K, water condensed in the upper atmosphere and water clouds appeared. This occurred within its first 100 million years. Within less than a gigayear, when Teff reached ∼160 K, ammonia clouds appeared on top of the water clouds, and this layering persists to this day. Stellar irradiation retards cloud formation, as does a large EGP mass, which keeps the EGP hot longer. Around a G2V star like the Sun, at 5 Gyr and for an EGP mass of 1.0 MJ , water clouds form at 1.5 AU, whereas ammonia clouds form beyond 4.5 AU [33]. Jupiter’s and Saturn’s current effective temperatures are 124.4 K and 95 K, respectively. Jupiter’s orbital distance and age are 5.2 AU and 4.6 Gyr. The orbital distance, mass, and radius of a coeval Saturn are 9.5 AU, 0.3 MJ , and 0.85 RJ . However, as an EGP of whatever mass cools, its atmospheric composition evolves through a similar chemical and condensation sequence. Figure 1 depicts the atmospheric temperature/pressure (T/P) profile for a sequence of 1- MJ , 5-Gyr models as a function of orbital distance from a G2V star. As the planet “moves” outward, its atmospheric temperature at a given pressure decreases. Superposed on the plot are the H2O and NH3 condensation lines. In an approximate sense, a given atmospheric composition and temperature can result from many combinations of orbital distance, planet mass, stellar type, and age. This lends an added degree of complexity to the study of EGPs with which the study of stars does not need to wrestle. The atmospheres of close-in EGPs (“roasters”) at orbital distances of ∼0.02-0.07 AU from a G, F, or K star are heated and maintained at temperatures of 1000-2000 K, roughly independent of planet mass or composition. An edge star of the solar-composition, hydrogen- burning main sequence (M∗∼75 MJ ) has a Teff of ∼1700 K. Therefore, an irradiated EGP, with a radius comparable to that of such a star, can be as luminous. Its atmospheric composition is predominantly H2, He, H2O, Na, K, and CO. At high temperatures, carbon is generally in carbon monoxide. This is the dominant molecule of carbon for M dwarfs with Teffs of 2200-3500 K. At the highest Teffs, clouds of iron particulates can form and 6 persist in the upper atmosphere, as may be the case in HD209458b. There are, however, significant day/night differences and unique reflective properties that distinguish a roaster from a lone and isolated edge star. Exotic general circulation models (GCMs) [34–36] may soon be necessary to understand the equatorial currents, jet streams, day/night differences, terminator chemistry, and global wind dynamics of severely irradiated roasters, in particular, and of orbiting, rotating EGPs, in general. It is useful to note that a young EGP in a wide orbit with a mass of 1.0 to 5.0 MJ has an atmosphere and spectrum that are similar to those of an old brown dwarf with a mass of 30-60 MJ . As it evolves, the spectroscopic class of a giant planet can transition from that of a hot M dwarf, into an L dwarf (where the silicate clouds are in the atmosphere), then into a T dwarf, ending up in the territory, as yet unexplored, between the Jovian planets and the “stars.” If its mass is low enough, an EGP can cool within gigayears to assume the aspect of our Jovian planets. Hence, by chemistry, clouds, and Teff, the study of brown dwarfs and EGPs are inextricably linked. Finally, the best theoretical fits to Saturn’s internal structure suggest that it contains a 5– 20 Earth-mass core of heavy elements [37]. This core of (perhaps) ice and rock may have been the nucleus around which Saturn formed and resembles the ice giants Neptune and Uranus. The latter may be aborted giant planets that were able to accrete but little hydrogen from the protosolar/protoplanetary nebula. The Neptune-mass extrasolar planet, 55 Cancri e, may be a stripped or aborted EGP. An alternative mode of giant planet formation is by direct collapse [25]. Under such a scenario, one would expect a closer correspondence between the heavy-element abundances of planet and parent star. Hence, the composition of its atmosphere and its heavy-element-dependent radius might be keys to an EGP’s formation. These are in principle measureable. —————————————————————————————————————— Box: Brown Dwarfs Brown dwarfs are substellar-mass objects (SMOs) (< ∼ 0.07 solar masses ≡∼75 7 MJ ) that are unable to ignite light hydrogen stably to become a star, but are otherwise formed like stars. The radiative surface losses of a star balance the thermonuclear power generated in its core. This requires sufficient mass. The surface losses of a less-massive brown dwarf are not fully compensated by ther- monuclear burning and it cools inexoribly after formation over a Hubble time. Nevertheless, brown dwarfs constitute the low-mass, low-temperature extension of the stellar family and are an important subject in their own right [29,38]. Masses in the range of ∼10 MJ to ∼75 MJ are frequently discussed, but overlap with the mass distribution of the EGP family is entirely possible. —————————————————————————————————————— IV. SPECTRAL FEATURES OF EGPS In principle, as with stellar atmospheres, direct detection of the spectrum of an extrasolar giant planet can reveal its elemental composition, radius, gravity (GMp/Rp 2), and Teff. Furthermore, when a cloud dwells in its atmosphere, its associated absorption and scattering properties might be used to determine the cloud’s particle size and makeup. Moreover, short- term temporal variations of the planet’s flux and spectrum might indicate rotation and/or meteorology. Finally, irradiation introduces the star-planet-Earth angle as an important parameter, so the orbit’s orientation and instantaneous orbital phase must be factored in (§V). Along with the dependences on Mp, age, stellar type, and orbital distance, this variety of influences and parameters makes the study of EGP spectral signatures and light curves, and their inversion to obtain planetary properties, rather complicated. Nevertheless, the molecular mix described in §III determines the emergent and reflected spectrum. Though H2 is abundant, it has no permanent electric dipole moment, and, hence, a very low photon absorption cross section in the optical and infrared. Similarly, helium is all but transparent. The result is that gaseous water vapor, with its strong absorption features from 0.94 µm to ∼7 µm , can define much of an EGP’s spectrum. Because water 8 resides in both the Earth’s and an EGP’s atmosphere, the water bands that bracket and determine the Earth’s photometric windows at ∼1.0 µm (Z), ∼1.25 µm (J), ∼1.65 µm (H), ∼2.2 µm (K), ∼3.45 µm (L′), and ∼4-5 µm (M), through which ground-based infrared as- tronomy is possible, are exactly the same windows in an EGP or brown dwarf atmosphere through which emergent flux can pour. Thus, and fortuitously for brown dwarf observations, the emission peaks for SMOs coincide with the classic Terrestrial atmospheric bandpasses. In lieu of measurements, theory fills the vacuum. Figure 2, taken from Burrows, Sudarsky, & Hubeny [33], depicts “phase-averaged” [39] planet/star flux ratios (f ) from 0.5 µm to 30 µm for a 1-MJ /5-Gyr EGP in a circular orbit at various distances from a G2V star like the Sun. These models are the same as those depicted in Fig. 1. Similar plots for different assumed parameters can be generated. The water absorption troughs are manifest throughout. For the closer EGPs at higher atmospheric temperatures, carbon resides in CO and methane features are weak. For these close-in EGPs, the Na-D line at 0.589 µm and the corresponding resonance line of K I at 0.77 µm are important absorbers, suppressing flux in the visible bands. Otherwise, the optical flux is buoyed by Rayleigh scattering of stellar light. As a increases, methane forms and the methane absorption features in the optical (most of the undulations seen in Fig. 2 for a > ∼ 0.5 AU shortward of 1 µm ), at ∼3.3 µm , and at ∼7.8 µm appear. Concomitantly, Na and K disappear from the atmosphere and the fluxes from ∼1.5 µm to ∼4 µm drop. For all models, the mid-infrared fluxes longward of ∼4 µm are due to self-emission, not reflection. As Fig. 2 makes clear, for larger orbital distances a bifurcation between a reflection component in the optical and an emission component in the mid-infrared appears. This separation into components is not so straightforward for the closer, more massive, or younger family members. For these EGPs, either the large residual heat coming from the core or the severe insolation prop up the fluxes from 1 to 4 µm . The more massive EGPs, or, for a given mass, the younger EGPs, have larger J, H, and K band fluxes. As a result, these bands are diagnostic of mass and age. For EGPs with large orbital distances, the wavelength range from 1.5 µm to 4 µm between the reflection and emission components may be the least favorable search space, unless the SMO is massive or young. 9 When water or ammonia clouds form, scattering off them enhances the optical fluxes, while absorption by them suppresses fluxes at longer wavelengths in, for example, the 4–5 µm window. Because water and ammonia clouds form in the middle of this distance sequence, the reflection efficiency (or “albedo”; §V) is not a monotonic function of a. These effects are incorporated into Fig. 2, but their precise magnitude depends upon unknown cloud particle size, composition, and patchiness. As a consequence, direct spectral measurements might constrain cloud properties. Importantly, trace non-equilibrium molecular species, difficult to model, can be present in quantities sufficient to alter colors. Such a “chromophore,” whose molecular nature is not yet known, absorbs in the blue and creates the reddish cast of both Jupiter and Saturn, lowering their albedos shortward of 0.55 µm by a factor of ∼1.5–2. (Chromophores were not modelled to produce Fig. 2.) As Fig. 2 suggests, the planet/star contrast ratio is better in the mid- to far-infrared, particularly at wide separations. For such separations, the contrast ratio in the optical can sink to 10−10. For the closest-in EGPs, such as HD209458b, OGLE-TR56b, 51 Peg b, and τ Boo b, the contrast ratio in the optical is between 10−5 and 10−6 and is more favorable. Such EGPs are not shown in Fig. 2; there are in fact about 20 known EGPs with orbital distances less than 0.08 AU. Due to the possible formation of iron clouds in their atmospheres, HD209458b and OGLE-TR56b may be brighter in the optical than 51 Peg b and may have higher reflection albedos. Figure 3 portrays a generic absolute flux spectrum at 10 parsecs of a close-in EGP (“Class V” in the nomenclature of Sudarsky et al. [39]), not unlike HD209458b. Highlighted are the positions of some of the important spectral features. Since modern telescopes can easily detect fluxes at the milliJansky level, Fig. 3 demonstrates that the fluxes themselves are not small. The problem is seeing the planet from under the glare of the star (§VI). At 10 parsecs and an orbital separation of 0.05 AU, the maximum angular separation is a challenging ∼5 milliarcsecs. 10 V. PHASE FUNCTIONS FOR EGPS AND ORBITAL ORIENTATION The planetary albedo and the phases executed as planets traverse their orbits are central quantities in the theory of EGP light curves. In addition, as we discuss in this section, the wavelength-dependent albedos are strong functions of orbital distance as well. Furthermore, the changing orientation of the illuminated face of a planet from the Earth’s perspective translates into a light curve that can show significant flux and color variations. In Fig. 2, these variations were averaged out over the orbit, assumed circular. The longitude indepen- dence of Jupiter’s T /P profile results in little day/night variation in the mid-infrared and, hence, little phase variation, but such a planet is too cold to be self-luminous enough in the optical for its reflected component not to dominate at these shorter wavelengths. Hence, Jupiter’s optical fluxes can vary from superior conjunction (full face) to first quarter (90◦ from superior conjunction, not seen from Earth) or last quarter (270◦ from superior con- junction, also not seen from Earth) by a factor of ∼3 [40]. Note that the phase dependence of an EGP’s light curve in the mid-infrared will depend on the degree to which heat can be efficiently redistributed over its entire face. This will depend on 3-dimensional GCM effects that have not been worked out. For the close-in EGPs, due to expected day/night temperature differences [41,42], it is likely that there will be phase variations at all wave- lengths. In particular, phase variations at thermal wavelengths are likely to shed light on the atmospheric dynamics and longitudinal temperature distribution of an EGP. Since its orbit and orientation play such an important role in an EGP’s flux at the Earth and in its interpretation, we summarize the basic formulae and concepts. We restrict ourselves to the optical, for which the concept of an albedo has a clear meaning, but note that the approach we summarize has general applicability. The planet/star flux ratio (f ) is given by: f = p(Rp/R)2Φ(α), (1) where R is the planet/star distance, p is the geometric albedo, Φ(α) is the phase function, and α is the star-EGP-Earth angle. Φ(α) is normalized to be 1.0 at full face, thereby defining 11 the geometric albedo, and is a decreasing function of α. For so-called “Lambert” reflection in which an incident ray on a planetary patch emerges uniformly over the exit hemisphere, p is 2/3 for purely scattering atmospheres and Φ(α) is given by the formula: Φ(α) = sin(α) + (π − α) cos(α) π . (2) However, EGP atmospheres are absorbing and the anisotropy of the single scattering phase function for grains, droplets, or molecules results in non-Lambertian behavior. For instance, back-scattering off cloud particles can introduce an “opposition” effect for which the planet appears “anomalously” bright at small αs. This spike might be a useful signature of cloud particle size. Moreover, the light scattered from EGPs is likely to be strongly polarized [43]. The degree of polarization as a function of wavelength and phase angle α can also be used to determine cloud properties. However, polarization will be rather more difficult to measure. Both p and Φ(α) are functions of wavelength, but the wavelength-dependence of p is the most severe. In fact, for cloud-free atmospheres, due to strong absorption by molecular bands, p can be as low as 0.03. Rayleigh scattering serves to support p, but mostly in the blue and UV, where, however, chromophores can decrease it. The presence of clouds increases p significantly. For instance, at 0.48 µm , Jupiter’s geometric albedo is ∼0.46 and Saturn’s is 0.39 [44]. Note that for orbital distances less than 1.5 AU, we expect the atmospheres of most EGPs to be clear. The albedo would be correspondingly low. As a consequence, the theoretical albedo is very non-monotonic with distance, ranging in the visible from perhaps ∼0.3 at 0.05 AU, to ∼0.05 at 0.2 AU, to ∼0.4 at 4 AU, to ∼0.7 at 15 AU [33,39,45]. In the visible (∼0.55 µm ), the geometric albedo for a roaster is severely suppressed by Na-D at 0.589 µm . Due to a methane feature, the geometric albedo can vary from 0.05 at ∼0.6 µm to ∼0.4 at 0.625 µm . Hence, variations with wavelength and with orbital distance by factors of 2 to 10 are not unexpected. Those planning programs of direct detection should be aware of such possibilities. Φ(α) and p must be calculated or measured, but the sole dependence of Φ(α) on α belies the complications introduced by an orbit’s inclination angle (i), eccentricity (e), argument 12 of periastron (ω), and longitude of ascending node (Ω). Along with the period (P ) and an arbitrary zero of time, these are the so-called Keplerian elements of an orbit. Figure 4 diagrams and defines these orientational and orbital parameters. In the plane of the orbit, the angle between the planet and the periastron/periapse (distance of closest approach to the star) at the star is θ. In the jargon of celestial mechanics, θ is the so-called “true anomaly.” For an edge-on orbit (i = 90◦), and one for which the line of nodes is perpendicular to the line of sight (Ω = 90◦) and parallel to the star-periapse line (ω = 0◦), θ is complementary to α (α = 90◦ − θ). As a result, θ = 0◦ at α = 90◦ (greatest elongation) and increases with time. Also, for such an edge-on orbit, α = 0◦ at superior conjunction. In general, cos(α) = sin(θ + ω) sin(i) sin(Ω) − cos(Ω) cos(θ + ω) . (3) This is merely an application of the law of cosines. For a circular orbit, R is equal to the semi-major axis (a). However, a planet in an eccentric orbit can experience significant variation in R, and, therefore, stellar insolation (by a factor of ( 1+e 1−e )2). For example, if e = 0.3, the stellar flux varies by ∼3.5 along its orbit. For e = 0.6, this variation is a factor of 16! Such eccentricities are by no means rare in the sample of known EGPs (cf. Table 1). Therefore, it is possible for the composition of an EGP atmosphere to change significantly during its orbit, for clouds to appear and disappear, and for there to be delays (“hysteresis”) in the accommodation of a planet’s atmosphere to a varying “insolation” regime. Ignoring the latter, eqs. 1 and 3 can be combined with Φ(α) and the standard Keplerian formula connecting θ and time for an orbit with a given P and e to derive an EGP’s light curve as a function of wavelength, i, e, Ω, ω, and time. The upshot is that, depending upon orientation and eccentricity, the brightness of an EGP can vary in its orbit not at all (for a face-on EGP in a circular orbit) or quite dramatically (e.g., for highly eccentric orbits at high inclination angles). Since astrometric measurements of stellar wobble induced by EGPs can yield the entire orbit (including inclination), data from the Space Interferometry Mission (SIM) [46] (expected to achieve 1-microarcsecond narrow-angle accuracy) or Gaia [47] could provide important supplementary data to aid in 13 the interpretation of direct detections of EGPs. As Saturn itself demonstrates, depending upon orbital orientation, planetary rings can greatly augment reflected light [40,48]. Their possible presence is a wild card in the interpre- tation of direct EGP signatures. Also, since Φ(α) is wavelength-dependent, the potentially large variation in reflected optical flux with epoch will be complemented by an interest- ing variation in color. The phase functions Φ(α) are wavelength-dependent. For example, planets should execute trajectories in the color-color space V − R vs. B − V , where B, V , and R are the standard blue, visible, and red bands. These trajectories will be functions of cloud particle size, among other things, and will be useful atmospheric diagnostics. Similar behavior in the near- and mid-infrared colors, though more modest, may be seen. VI. GROUND-BASED AND SPACE-BASED TELESCOPES FOR DIRECT DETECTION As Fig. 2 implies, the wide range of planet/star contrast ratios and spectral diagnostics suggests different technological solutions to direct detection. Furthermore, the relative mer- its of searching in the optical, near-infrared, or the mid-infrared have yet to be determined. Both ground-based (less expensive) and space-based (more capable) paths are being pursued and while a discussion that does justice to the many initiatives whose goal is the remote sensing of EGPs is far beyond the scope of this review, we summarize a few representative approaches. EGPs, especially if they are young, massive, and close, are bright enough that current 8- to 10-meter class ground-based telescopes or near-term space telescopes (such as the James Webb Space Telescope (JWST) [49] with a 6-meter aperture) might be sensitive enough to pick up their light. This is particularly true in the near- and mid-infrared (see Figs. 2 and 3). However, under the extreme glare of its parent star and at small angular separations of from ∼milliarcsecs to around an arcsecond (Table 1), traditional telescope optics spills far too much light in the vicinity of the planet. The major culprits on the ground are the 14 turbulence of the atmosphere (“seeing” and scintillation), scattering off dust and the spider mount of the secondary, and imperfections in the mirror(s). Also at issue is the stability of the optical system. Even for perfect optics, the diffraction pattern due to the finite telescope aperture leaves a characteristic “Airy” pattern that for a “Jupiter” at 10 parsecs around a solar-type star would in the optical be hundreds of times brighter. Hence, for large (8–10-meter) ground-based telescopes, (e.g., the two Kecks [50,51], the four VLTs [52,53], the two Geminis [54], Subaru [55], the binocular LBT [56,57]) and mam- moth proposed telescopes (e.g., the 100-meter OWL [58], the 20-meter GMT [59], the 30- meter GSMT [60]), special efforts will be required. These include adaptive optics (AO) to compensate for atmospheric fluctuations (and mirror imperfections) with many hundreds or thousands of fast (millisecond) actuators and very accurate wavefront sensing. The latter can, in principle, be achieved using artificial laser guide stars or stars in the field of view. (With AO, there is usually a bright star close enough to obviate the need for an artificial beacon.) Interferometry to null out the stellar light is also being pursued by the LBT, VLT, and Keck, and the depth of the null is crucial, as is the angular region over which a sufficient null can be achieved. Finally, apodizing masks and/or coronagraphic spots to occult the star and, by diffractive interference, redistribute the star’s light away from the planet are highly desirable (and may be necessary). Note that since Dome C in Antartica has some of the best seeing on the planet and the quietest atmosphere, placing a giant next-generation telescope there may have its advantages [61]. An EGP imaging system will be judged by the planet/star contrast ratio, f , it can achieve at a given wavelength and for a given angular separation from the star. Angles of 0.05 to 2.0 arcsecs are contemplated (Table 1), with the requisite contrasts at smaller angles deemed too difficult for first-generation imaging. Note that the actual requirements are a function of distance to the star. As Fig. 2 indicates for a 5-Gyr/1-MJ EGP at 10 parsecs, f s better than 10−4 at 10 µm and better than 10−8 in the optical might be necessary. At the 4-5 µm bump, f s from 10−5 to 10−8 may be called for. Fortunately, these performance goals can be relaxed for more massive and younger EGPs at angular separations greater than 15 ∼0.1′′. For the roasters, almost independent of mass and age, f reaches 10−3 in the mid- infrared and < ∼ 10−5 in the optical, but at the corresponding milliarcsecond separations even these contrast ratios may be too challenging for imaging. Figure 5 compares the theoretically required contrast ratios for the fiducial 5-Gyr/1-MJ EGP at various wavelengths (taken from Fig. 2) as a function of angular separation from a solar-type star at 10 parsecs with the putative capabilities of a sample of proposed imaging systems, both on the ground and in space. Contrast ratios for a 0.5-Gyr/7-MJ EGP in the H band are also shown. Orbit and orientation effects have been ignored and large error bars should be assigned to both theory and projected capability. In addition, care should be taken to compare theoretical numbers with experimental hopes for the same wavebands. In the interests of brevity, we have included multiple wavebands on Fig. 5. Telescopes are stages of components (primary mirror, secondary mirror, lens, apertures, apodizing masks, coronagraphs, etc.) that in series act on the incident source wavefront to focus light of a desired character on instruments. Daisy-chained together, each “optical” component convolves itself with an input wavefront to produce an output wavefront. In spatial frequency space, the operation of each stage along the optical path is to multiply the Fourier transform of the incident wavefront by a Fourier transform characteristic of that component’s optical properties and geometry. If you can introduce components in the optical path that filter or alter the frequency distribution of the wavefront in such a way that its inverse (the spatial distribution of the light in the last image plane) has little or no light in a 2-dimensional angular realm around the star where a planet might reside, then you have a planet-imaging system. Though a telescope’s classical angular resolution (∼λ/D, where D is the diameter of the telescope primary) improves with decreasing λ, the negative effects of the atmosphere actually diminish with increasing λ. As a result, many ground-based planet-finding initiatives (e.g., MMT(AO) [57], LBT-I, VLT-I, VLT-PF, Keck-I, Gemini- XAOPI/ExAOC, OWL) are planning to optimize in the near- or mid-infrared. Figure 5 summarizes the performance goals of some of them. In coronagraphic mode, the space-bourne JWST may achieve f s of 10−5 to 10−6 for 16 wavelengths from ∼1.0 µm to ∼5.0 µm . HST/NICMOS has already achieved comparable f s in H band at angular separations from 0.3′′ to 1.0′′ [62]. However, for single space telescopes without the atmosphere with which to contend, the optical is clearly preferred (small λ/D). Curiously, above the atmosphere mirror imperfections and thermal flexure are still problems and an AO system is necessary to cancel the wavefront errors introduced by the corrugations that remain on an otherwise almost perfect mirror surface after state-of- the-art machining and polishing. Two major space-based projects to image EGPs are being proposed. The first, EP IC [63], is a nulling coronagraph which converts a single telescope pupil into a multi-beam nulling interferometer, producing a null which is then filtered by an array of single-mode fibers to suppress the residual scattered light. The design goal of EP IC is for f s of 10−9 to 10−10. The second, ECLIPSE [64,65], is an off-axis coronagraph with an exquisitely-figured 1.8-meter primary that is designed to achieve f s in the V band better than 10−9 for angular separations from ∼0.1′′ to ∼2.0′′. Both ECLIPSE and EP IC will be challenging, but if successful will directly detect within ∼7 years many EGPs in the solar neighborhood out to 10–15 parsecs. However, the flagship of the NASA Origins program, the Terrestrial Planet Finder (TPF [66]), whose goal is to image planetary systems and extrasolar Earths and to obtain low- resolution (λ/∆λ = 10-20) spectra that may reveal in rudimentary fashion the O2, H2O, CH4, O3, or CO2 signatures of life, will also be a formidable instrument for directly detecting and characterizing EGPs. As Fig. 5 indicates, for angular separations between ∼0.05′′ and ∼2.0′′, either the more-straightforward optical coronagraphic design (TPF-C) or the multi- telescope infrared (5–20 µm ) interferometer (TPF-I/Darwin) would detect EGPs much more readily than the extrasolar Earths that are its primary targets. For the close-in EGPs, direct imaging seems out of the question for the forseeable future. However, this does not mean that the planetary flux can not be measured. From the ground, there are a variety of techniques to use the planet-plus-star light to distinguish the planetary component (particularly for known roasters). These include 1) using precision photometry to a part in 105 (!) to measure the phase variations of the summed optical or near-infrared light, 17 2) measuring the motion of the light centroid, perhaps best done in the mid-infrared with an Antarctic 30-meter telescope, 3) spectral deconvolution of a known EGP/star system using its RV-measured velocities and ephemeris, and 4) multi-frequency differential interferometric imaging (pioneered for Keck, among others). Further transit studies of HD209458b (such as led to the discovery of the Na-D and Lyman-α features) are certainly warranted. The ground- based methods will be challenging, but less expensive than space-based efforts. However, there is currently in space a micro-satellite, MOST [67], with a 15-centimeter aperture, that is designed to achieve photometric accuracy in the optical of a few×10−6. It has on its current observing manifest programs to stare at 51 Peg b, τ Boo b, and HD209458b. There have as yet been no announcements. VII. THE FUTURE With many programs of direct planet detection planned on a large subset of the LBT, VLT, Keck, Gemini, GMT, GSMT/CELT, and OWL on the ground and HST, Corot, Kepler, MOST, SIM, Gaia, TPF-C, TPF-I/Darwin, ECLIPSE, EPIC, and JWST in space, during the next twenty years there will be an increasing crescendo of new results on extrasolar planets that will completely transform our view of the nature of planetary systems. EGPs, being brighter, are the natural technological and scientific stepping stones on the path to imaging extrasolar Earths. We will encounter them first. Both the NASA and ESA roadmaps [68] have given planet detection pride of place. Strategies are now being formulated to establish a logical sequence of missions and telescope construction that will optimize the pace of discovery. Moreover, theoretical work in support of mission planning is maturing to the point that it may be ready to interpret what we observe. However, a theorist’s prejudices aside, one can’t help but wonder: What is it we will actually find? What discoveries will be made? As the hunt for worlds beyond our solar system quickens, an ancient curiosity stirs to ask: What will our generation see from that fabled peak in Darien [69]? 18 REFERENCES [1] Mayor, M. & Queloz, D. A Jupiter-Mass Companion to a Solar-Type Star. Nature 378, 355 (1995). [2] Marcy, G.W. and R.P. Butler, R.P. A Planetary Companion to 70 Virginis. Astrophys. J. 464, L147-151 (1996). [3] see J. Schneider’s Extrasolar Planet Encyclopaedia at http://www.obspm.fr/encycl/encycl.html for an up-to-date listing. [4] MacArthur, B. et al. Detection of a Neptune-mass planet in the ρ1 Cancri system using the Hobby-Eberly Telescope. (55 Cancri e) submitted to Astrophys. J. (2004). [5] Konacki, M., Torres, G., Jha, S., and Sasselov, D. An extrasolar planet that transits the disk of its parent star. Nature 421, 507-509 (2003). [6] Torres, G, Konacki, M., Sasselov, D., & Jha, S. New Data and Improved Parameters for the Extrasolar Transiting Planet OGLE-TR-56b. Astrophys. J. 609, 1071-1075 (2003). [7] Sasselov, D. The New Transiting Planet OGLE-TR-56b: Orbit and Atmosphere. Astro- phys. J. 596, 1327-1331 (2003). [8] Henry, G., Marcy, G.W., Butler, R.P., & Vogt, S.S. A Transiting “51 Peg-like” Planet. Astrophys. J. 529, L41-44 (2000). [9] Charbonneau, D., Brown, T.M., Latham, D.W., & Mayor, M. Detection of Planetary Transits Across a Sun-like Star. Astrophys. J. Letters 529, L45-48 (2000). [10] Brown, T. M., Charbonneau, D., Gilliland, R.L., Noyes, R.W., and Burrows, A. Hub- ble Space Telescope Time-Series Photometry of the Transiting Planet of HD 209458. Astrophys. J. 552, 699-709 (2001). [11] Bouchy, F., Pont, F., Santos, F.C., Melo, C., Mayor, M., Queloz, D., & Udry, S. Two 19 new “very hot Jupiters” among the OGLE transiting candidates. Astron. Astrophys. 421, L13-16 (2004). [12] Alonso, R. et al. TrES-1: The transiting planet of a bright K0V star. in press (2004). [13] Pont, F., Bouchy, F., Queloz, D., Santos, N.C., Melo, C., Mayor, M., & Udry, S. The “missing link”: A 4-day period transiting exoplanet around OGLE-TR-111. Astron. Astrophys. , in press (astro-ph/0408499) (2004). [14] Koch, D., Borucki, W., Webster, L., Dunham, E., Jenkins, J., Marrion, J., & Reitsema, H. Kepler: a space mission to detect earth-class exoplanets. SPIE Conference 3356: Space Telescopes and Instruments V, 599-607 (1998). [15] Antonello, E. & Ruiz, S.M. The Corot Mission. Memoria della Societa Astronomica Italiana 73, 1241 (2002). [16] Burrows, A., Sudarsky, D., & Hubbard, W.B. A Theory for the Radius of the Transiting Giant Planet HD 209458b. Astrophys. J. 594, 545-551 (2003). [17] Burrows, A., Hubeny, I., Hubbard, W.B., & Sudarsky, D. Theoretical Radii of Transiting Giant Planets: The Case of OGLE-TR-56b. Astrophys. J. 610, L53-56 (2004). [18] Baraffe, I., Chabrier, G., Allard, F., and Hauschildt, P.H. Evolutionary models for cool brown dwarfs and extrasolar giant planets. The case of HD 209458. Astron. Astrophys. 402, 701-712 (2003). [19] Charbonneau, D., Brown, T.M., Noyes, R.W., & Gilliland, R.L. Detection of an Extra- solar Planet Atmosphere. Astrophys. J. 568, 377-384 (2002). [20] Seager, S. & Sasselov, D.D. Theoretical Transmission Spectra during Extrasolar Giant Planet Transits. Astrophys. J. 537, 916-921 (2000). [21] Hubbard, W.B., Fortney, J.F., Lunine, J.I., Burrows, A., Sudarsky, D., & Pinto, P.A. Theory of Extrasolar Giant Planet Transits. Astrophys. J. 560, 413-419 (2001). 20 [22] Vidal-Madjar, A., des Etangs, A., Desert, J.-M., Ballester, G.E., Ferlet, R., Hebrard, G., & Mayor, M. An extended upper atmosphere around the extrasolar planet HD209458b. Nature 422, 143-146 (2003). [23] Burrows, A. & Lunine, J.I. Extrasolar Planets - Astronomical Questions of Origin and Survival. Nature 378, p. 333 (1995). [24] Mizuno, H. Formation of the Giant Planets. Progress of Theoretical Physics 64, 544-557 (1980). [25] Boss, A.P. Gas Giant Protoplanet Formation: Disk Instability Models with Thermody- namics and Radiative Transfer. Astrophys. J. 563, 367-373 (2001). [26] Trilling, D., Benz, W., Guillot, T., W.B. Hubbard, W.B., J.I. Lunine, & Burrows, A. Orbital Migration of Giant Planets: Modeling Extrasolar Planets. Astrophys. J. 500, 428-439 (1998). [27] Fortney, J.J. & Hubbard, W.B. Effects of Helium Phase Separation on the Evolution of Extrasolar Giant Planets. Astrophys. J. 608, 1039-1049 (2004). [28] Burgasser, A. et al. The Spectra of T Dwarfs. I. Near-Infrared Data and Spectral Clas- sification. Astrophys. J. 564, 421-451 (2002). [29] Burrows, A., Hubbard, W.B., Lunine, J.I., & Liebert, J. The theory of brown dwarfs and extrasolar giant planets. Rev. Mod. Phys. 73, 719-765 (2001). [30] Burrows, A. & Sharp, C.M. Chemical Equilibrium Abundances in Brown Dwarf and Extrasolar Giant Planet Atmospheres. Astrophys. J. 512, 843-863 (1999). [31] Lodders, K. Alkali Element Chemistry in Cool Dwarf Atmospheres. Astrophys. J. 519, 793-801 (1999). [32] Lodders, K. & Fegley, B. Atmospheric Chemistry in Giant Planets, Brown Dwarfs, and Low-Mass Dwarf Stars. I. Carbon, Nitrogen, and Oxygen. Icarus 155, 393-424 (2002). 21 [33] Burrows, A., Sudarsky, D., & Hubeny, I. Spectra and Diagnostics for the Direct Detec- tion of Wide-Separation Extrasolar Giant Planets. Astrophys. J. 609, 407-416 (2004). [34] Menou, K., Cho, J. Y-K., Seager, S. & Hansen, B.M.S. “Weather” Variability of Close-in Extrasolar Giant Planets. Astrophys. J. 587, L113-116 (2003). [35] Cho, J. Y-K., Menou, K., Hansen, B.M.S., & Seager, S. The Changing Face of the Extrasolar Giant Planet HD 209458b. Astrophys. J. 587, L117-120 (2003). [36] Burkert, A., Lin, D.N.C., Bodenheimer, P., Jones, C., & Yorke, H. On the Surface Heat- ing of Synchronously-Spinning Short-Period Jovian Planets. astro-ph/0312476 (2004). [37] Saumon, D. & Guillot, T. Shock Compression of Deuterium and the Interiors of Jupiter and Saturn. Astrophys. J. 609, 1170-1180 (2004). [38] Burrows, A., Marley, M., Hubbard, W.B., Lunine, J.I., Guillot, T., Saumon, D., Freed- man, R., Sudarsky, D., & Sharp, C. A Nongray Theory of Extrasolar Giant Planets and Brown Dwarfs. Astrophys. J. 491, 856-875 (1997). [39] Sudarsky, D., Burrows, A., & Pinto, P. Albedo and Reflection Spectra of Extrasolar Giant Planets. Astrophys. J. 538, 885-903 (2000). [40] Dyudina, U.A., Sackett, P.D., Bayliss, D.D.R., Seager, S., Porco, C., Throop, H.B., & Dones, L. Phase light curves for extrasolar Jupiters and Saturns. astro-ph/0406390 (2004). [41] Guillot, T. & Showman, A.P. Evolution of “51 Pegasus b-like” planets. Astron. Astro- phys. 385, 156-165 (2002). [42] Showman, A.P. & Guillot, T. Atmospheric circulation and tides of “51 Pegasus b-like” planets. Astron. Astrophys. 385, 166-180 (2002). [43] Seager, S., Whitney, B.A, & Sasselov, D.D. Photometric Light Curves and Polarization of Close-in Extrasolar Giant Planets. Astrophys. J. 540, 504-520 (2000). 22 [44] Karkoschka, E. Methane, Ammonia, and Temperature Measurements of the Jovian Planets and Titan from CCD-Spectrophotometry. Icarus 133, 134-146 (1999). [45] Sudarsky, D., Burrows, A., & Hubeny, I. Theoretical Spectra and Atmospheres of Ex- trasolar Giant Planets. Astrophys. J. 588, 1121-1148 (2003). [46] Unwin, S.C. & Shao, M. Space Interferometry Mission. in Interferometry in Optical Astronomy (eds. P. J. Lena & A. Quirrenbach) 754-761 (2000). [47] Perryman, M.A.C. GAIA Spectroscopy: Science and Technology. in ASP Conference Proceedings, Vol. 298 (ed. Ulisse Munari) p. 3 (ASP Conf. Series, 2003). [48] Arnold, I. & Schneider, J. The detectability of extrasolar planet surroundings I. Reflected-light photometry of unresolved rings. astro-ph/0403330 (2004). [49] Mather, J.C. and Stockman, H.S. Proc. SPIE 4013, 2-16 (2000). [50] Akeson, R.L., Swain, M. R., & Colavita, M.M. Differential phase technique with the Keck Interferometer. in Interferometry in Optical Astronomy (ed. P. J. Lena) Proc. SPIE 4006, 321-327 (2000). [51] Akeson, R.L. & Swain, M.R. Differential Phase Observations of Extrasolar Planets with the Keck Interferometer. in From Giant Planets to Cool Stars, (ed. C.A. Griffith & M.S. Marley) p. 300 (ASP Conference Series Vol. 212, 2000) [52] Paresce, F. Scientific Objectives of the VLTI Interferometer. The Messenger 104, 5-7 (2001). [53] Beuzit, J.-L., et al. The Planet Finder project for the VLT. in SF2A-2004: Semaine de l’Astrophysique Francaise (eds. F. Combes, D. Barret, T. Contini, F. Meynadier & L. Paganii) p. 32 (EdP-Sciences, Conference Series, 2004). [54] Macintosh, B. et al. Extreme Adaptive Optics Planet Imager (XAOPI). in Techniques and Instrumentation for Detection of Exoplanets. (ed. Coulter, D. R.) pp. 272-282 (Pro- 23 ceedings of the SPIE, Vol. 5170, 2003). [55] Tamura, Itoh, I., & Oasa, Y. Searches for Extrasolar Planets with the Subaru Telescope: Companions and Free-Floaters. in Proceedings of the IAU 8th Asian-Pacific Regional Meeting: Volume I, (ed. Satoru Ikeuchi, John Hearnshaw and Tomoyuki Hanawa) pp. 73-76 (San Francisco, Astronomical Society of the Pacific, Vol. 289, 2003). [56] Hinz, P.M., Angel, J.R.P., Woolf, N.J., Hoffman, W.F., & McCarthy, D.W. BLINC: a testbed for nulling interferometry in the thermal infrared. in Interferometry in Optical Astronomy (ed. P.J. Lena) Proc. SPIE 4006, 349 (2000). [57] Hinz, P.M. Nulling interferometry for studying other planetary systems: Techniques and observations. PhD Thesis, The University of Arizona (2001). [58] Hawarden, T.G., Gilmozzi, R., & Hainaut, O. Using a 100-meter ELT (e.g., “OWL”) for Extrasolar Planet and Extrasolar Life Detection. in Scientific Frontiers in Research on Extrasolar Planets ASP Conference Series, Vol. 294 (ed. Drake Deming and Sara Seager) pp. 581-586 (San Francisco: ASP, 2003). [59] Davison, W.B., Woolf, N.J., & Angel, J.R.P. Design and Analysis of 20-m track mounted and 30-m telescopes. in Future Giant Telescopes (eds. Angel, J.R.P. & Gilmozzi, R.) Proceedings of the SPIE, Vol. 4840, pp. 533-540 (2003). [60] Strom, S. Toward a Giant Segmented Mirror Telescope: A Progress Report from AURA’s New Initiatives Office. American Astronomical Society Meeting 203, # 24.04 (2003). [61] Aristidi, E., Agabi, A., Vernin, J., Azouit, M., Martin, F., Ziad, A., & Fossat, E. Antarctic site testing: First daytime seeing monitoring at Dome C. Astron. Astrophys. 406, L19-L22 (2003). [62] Schneider, G. et al. Domains of observability in the near-infrared with HST/NICMOS and adaptive-optics augmented large ground-based telescopes. unpublished manuscript 24 at http://nicmosis.as.arizona.edu:8000/REPORTS/NICMOS_AO_WHITEPAPER.html (2002). [63] Shao, M., Levine, B.M., Wallace, J.K., Serabyn, E., & Liu, D.T. The Visible Nulling Coronagraph–Progress Towards Mission and Technology Development. BAAS 203, #03.04 (2003). [64] Trauger, J. et al. Eclipse, A Direct Imaging Investigation of Nearby Planetary Systems. BAAS 197, # 49.07, p. 1486 (2000). [65] Trauger, J., Hull, A.B., & Redding, D.A. Eclipse Test Bed for Very High Contrast Space Astronomy. BAAS 199, # 86.04 p. 1431 (2001). [66] Beichman, C.A., Coulter, D.R., Lindensmith, C., & Lawson, P.R. Selected Mission Ar- chitectures For The Terrestrial Planet Finder (TPF): Large, Medium, and Small. in Future Research Direction and Visions for Astronomy. (ed. Dressler, Alan M.), Pro- ceedings of the SPIE, Vol. 4835, pp. 115-121 (2002). [67] Matthews, J.M., Kuschnig, R., & Shkolnik, E. Ultraprecise photometry from space: exploring pulsations with the “Humble Space Telescope.” in Proceedings of the SOHO 10/GONG 2000 Workshop: Helio- and asteroseismology at the dawn of the millen- nium (ed. A. Wilson) ESA SP-464, Noordwijk: ESA Publications Division, pp. 385-390 (2001). [68] Roadmap for the Office of Science Origins theme. NASA/JPL-400-1060 (2003); NASA 2003 Strategic Plan. Document # NP-2003-01-298-HQ (2003). [69] Keats, J. On First Looking into Chapman’s Homer. (1816). [70] Racine, R., Walker, G.A.H., Nadeau, D., Doyon, R. & Marois, C. Speckle Noise and the Detection of Faint Companions. Publications of the Astronomical Society of the Pacific 111, 587-594 (1999). Correspondence and requests for materials should be addressed to Adam Burrows (bur- 25 [email protected]). The author acknowledges David Sudarsky, Ivan Hubeny, Bill Hubbard, Jonathan Lunine, Jim Liebert, Jason Young, John Trauger, Jonathan Fortney, Aigen Li, Christopher Sharp, and Drew Milsom for fruitful conversations or technical aid and help during the course of this work, as well as NASA and the NASA Astrobiology Institute for their financial support. 26 TABLES TABLE I. Interesting EGPs Listed by Angular Separation EGP a(1 + e)/d (′′)a star a (AU) d (pc) P Mpsin(i) (MJ ) e 3.2 6.85 yrs. 0.86 4.05 0.76 ∼1 4.61 10.3 1.15 3.3 2.54 7.71 1.17 1.25 1.0 2.1 1.7 2.0 4.9 6.81 17.1 6.59 2.11 1.89 0.56 0.61 0.16 0.1 ∼0.8 0.41 0.62 ∼0 0.33 0.06 0.41 0.32 ∼0 0.52 0.34 0.31 0.10 0.42 0.54 0.23 0.12 0.18 0.1 0.27 2.38 31.03 2.61 37.44 13.5 241 days 4.72 4.72 61.0 30.1 ǫ Eri b 55 Cnc d 47 UMa c HD 160691c υ And d HD 39091b Gl 777A b 14 Her b 47 UMa b HD 33636b HD 10647b γ Cephei b HD 147513b HD 216437b HD 160691b HD 70642b HD 50554b HD 106252b HD 168443c HD 10697b υ And c GJ 876b GJ 876c 1.61 0.51 0.31 0.27 0.27 0.26 0.23 0.20 0.17 0.17 0.16 0.15 0.15 0.134 0.127 0.121 0.109 0.108 0.107 0.075 0.072 0.049 0.036 13.4 13.3 15.3 13.5 20.6 15.9 17 13.3 28.7 17.3 11.8 12.9 26.5 15.3 29 33 30 14.7 7.10 3.56 3.47 5.70 7.15 4.51 2.98 4.43 2.89 2.5 1.48 3.54 1.74 4.79 3.50 4.11 4.76 2.99 K2V G8V G0V 3.3 5.9 3.73 G3IV-V 2.3 F8V G1IV G6V K0V G0V G0V F9V K2V G3V G4V 2.50 3.34 3.65 2.5 2.09 3.56 2.10 1.8 1.26 2.7 G3IV-V 1.48 G5IV-V 3.3 F8V G0V G5V G5IV F8V M4V M4V 2.87 2.0 0.83 0.21 0.13 27 HD 114762b 0.017 55 Cnc b 8.4 × 10−3 F9V G8V 0.35 0.12 28 13.4 υ And b 4.5 × 10−3 F8V 0.059 13.5 51 Peg b 3.4 × 10−3 τ Boo b 3.3 × 10−3 G2V F7V 0.05 0.05 14.7 15 HD 49674b 1.6 × 10−3 G5V 0.057 40.7 HD 209458b 9.6 × 10−4 G0V 0.045 47 HD 83443b 9.4 × 10−4 K0V 0.038 43.5 OGLE-TR56b 1.5 × 10−5 G0V 0.023 ∼1500 84.0 14.7 4.62 4.23 3.31 4.95 3.52 2.99 1.21 11.0 0.84 0.71 0.44 4.09 0.12 0.69 0.35 1.45 0.34 0.02 0.034 0.01 ∼0 0.17 ∼0 0.08 ∼0 a Maximum possible angular separation (at apoapse/apastron). 28 Figure 1. Profiles of atmospheric temperature (in Kelvin) versus the logarithm base ten of the pressure (in bars) for a family of irradiated 1-MJ EGPs around a G2V star as a function of orbital distance. Note that the pressure is decreasing along the ordinate, which thereby resembles altitude. The orbits are assumed to be circular, the planets are assumed to have a radius of 1 RJ , and the orbital separations vary from 0.2 AU to 15 AU. The intercepts with the dashed lines identified with either {NH3} or {H2O} denote the positions where the corresponding clouds form. Taken from Burrows, Sudarsky, and Hubeny [33]. See text for a discussion. Figure 2. Planet to star flux ratios versus wavelength (in microns) from 0.5 µm to 30 µm for a 1-MJ EGP with an age of 5 Gyr orbiting a G2V main sequence star similar to the Sun. This figure portrays ratio spectra as a function of orbital distance from 0.2 AU to 15 AU. Zero eccentricity has been assumed and the planet spectra have been phase-averaged as described in Sudarsky, Burrows, and Hubeny [45]. The associated T /P profiles are given in Fig. 1. Note that the planet/star flux ratio is most favorable in the mid-infrared. Water features at 0.94 µm , 1.2 µm , 1.4 µm , 1.9 µm , 2.6 µm , and 6.5–8 µm , methane features in the optical and at 0.89 µm , 2.2 µm , 3.3 µm , and 7.8 µm , carbon monoxide features at 2.3 µm and 4.67 µm , the Na-D doublet at 0.589 µm , and the K I doublet at 0.77 µm help shape these spectra. Taken from Burrows, Sudarsky, and Hubeny [33]. See text for discussion. Figure 3. The logarithm of the absolute flux in milliJanskays (≡ 10−26 ergs cm−2 s−1 Hz−1) at 10 parsecs for a “Class V” roaster versus wavelength (in microns) from 0.4 µm to 5 µm . This could be a 1-MJ EGP in a 0.05 AU orbit around a solar-type star. The planet spectrum has been phase-averaged as described in Sudarsky et al. [39,45]. Shown are the positions of various relevant molecular bands and atomic lines. Figure taken from Sudarsky, Burrows, & Hubeny [45]. See text for discussion. Figure 4. Keplerian orbital elements. The intersection of the orbit plane with the observational plane defines the angles i, ω, Ω, and θ. The angle between the observer (Earth) and the line of nodes (intersection of the orbit plane with the horizontal plane) is the longitude of the ascending node (Ω), the angle between the line of nodes and the focus(star, 29 in yellow)–periapse(black dot) is the argument of periastron (ω), the angle between the orbit plane and the Z-axis (perpendicular to the horizontal plane) is the inclination (i), and the angle between the focus–periapse line and the position of the planet (red dot) is the true anomaly (θ). See text for details. Figure 5. A comparison of the planet/star contrast ratios (and contrast magnitudes = −2.5 log(f )) versus angular separation (in arcseconds) achievable for some proposed planet imaging systems. A distance of 10 parsecs is assumed. Integration times and signals-to-noise assumed vary and are taken from preliminary studies by the associated instrument teams. At H band (red), the imaging telescopes represented include the Canada/France/Hawaii Telescope (CFHT) [70], HST/NICMOS, and the Gemini/XAOPI. Not shown on this plot is the MMT (AO) system, which should achieve at a wavelength of 5 µm f s from 10−4 to a few ×10−6 for angular separations of 0.3′′ to 1.0′′, respectively. The LBT, also not shown, should achieve at a wavelength of 10 µm approximately 10× better performance than this. At 5 µm , a notional curve for a 20- or 30-meter telescope in Antarctica and the JWST (in fact at 4.6 µm ) are provided. At 10 µm , a notional curve for a 100-meter in Antarctica is given. Also included on this plot is the interferometric version of TPF (TPF-I/Darwin), which might have a sensitivity of one part in 107 from 5 µm to 20 µm . All the mid-infrared curves are in blue. In the optical (green, V ), putative sensitivities for EP IC, ECLIPSE, and TPF-C are plotted. Superposed are corresponding “phase-averaged” theoretical curves (dashed) for a 5-Gyr/1-MJ EGP around a G2V star in the H band (∼1.65 µm ), in the 4–5 µm band, and at 10 µm (see Fig. 2). Also included are a theoretical curve in the H band (dashed red) for a 0.5-Gyr/7-MJ EGP around a G2V star and a green swathe where the known EGPs may reside in the optical (V band). Note that the theoretical curves for more massive and younger EGPs than represented on this plot can be considerably higher. Orientation effects have been ignored. Each curve is for a given wavelength or bandpass and the imager and theory must be compared at the same wavelength. Very generous error bars should be assumed. The photometric sensitivity curves for MOST and Kepler are also superposed. See text for details. 30 -4 -2 0 0 500 1000 1500 -4 -6 -8 -10 -12 -14 -16 -18 1 10 This figure "figure5.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/astro-ph/0501484v1 z i Planet θ Periapse ω Ω Earth
astro-ph/0208022
1
0208
2002-08-01T06:38:52
GMRT radio observations of microquasar GRS 1915+105
[ "astro-ph" ]
We have observed microquasar GRS 1915+105 at 1.28 GHz for 8 days from June 18 to July 1, 2001 using Giant Metrewave Radio Telescope (GMRT). We have seen several isolated radio baby flares of varying intensity and duration. We have also observed broad composite flares with rise and decay times of few hours on June 28-29, 2001 few days prior to when source went to the ``plateau radio state'' on 3rd July, 2001. These broad radio flares consist of several overlapping baby radio flares. The source was in the low-hard X-ray state during this period. We compare these results with 15 GHz radio data from the Ryle telescope.
astro-ph
astro-ph
GMRT radio observations of microquasar GRS 1915+105 J. S. Yadav1, C. H. Ishwara-Chandra2, A. Pramesh Rao2 and G. G. Pooley3 1 TIFR, Homi Bhabha Road, Mumbai - 400 005 INDIA 2 NCRA, Post Bag No. 3, Ganeshkhind, Pune - 411 007 INDIA 3 Mullard Radio Astronomy Observatory, Cambridge, UK Abstract. We have observed microquasar GRS 1915+105 at 1.28 GHz for 8 days from June 18 to July 1, 2001 using Giant Metrewave Radio Telescope (GMRT). We have seen several isolated radio baby flares of varying intensity and duration. We have also observed broad composite flares with rise and decay times of few hours on June 28-29, 2001 few days prior to when source went to the "plateau radio state" on 3rd July, 2001. These broad radio flares consist of several overlapping baby radio flares. The source was in the low-hard X-ray state during this period. We compare these results with 15 GHz radio data from the Ryle telescope. 1. Introduction The long term monitoring of microquasar GRS 1915+105 has shown broad corre- lation between the X-ray and non-thermal radio emission [5]. The radio emission in GRS 1915+105 can be classified into three classes; (i) the relativistic superlu- minal radio jets of flux density ∼ 1 Jy with decay time-scales of several days ([4], [3]), (ii) the baby jets of 20 − 40 min durations with flux density of 20 − 200 mJy both in infrared (IR) and radio ([9], [2]), and (iii) the plateau state with persis- tent radio emission of 20 − 100 mJy for extended durations [8]. In the case of superluminal jets, the radio emission has steep spectra and are observed at large distances (400 − 5000 AU) from the accretion disk [3], [1]. The radio emission during other two classes has flat spectra and they occur close to the accretion disk (within a few tens of AU)[1]. The multi-wavelength studies have indicated strong disk-jet connection for class 2([7], [2], [11]). In this paper, we present our radio observations at 1.28 and 15 GHz during June 18 to July 3, 2001 (just prior to when source went to the plateau state on July 3). These observations include faint radio emission, isolated baby radio flares riding above smooth as well as slowly rising/decaying radio emission (broad flares), and large radio flare (∼ 200 mJy on June 20). 2. Observations and discussion The 1.28 GHz radio observations were carried out with a bandwidth of 16 MHz using the GMRT at Pune, India [10] on 2001 June 18, 22, 23, 27, 28, 29, 30 and July 1. The flux density scale is set by observing primary calibrator 3C286 or 3C48. The integration time of 32 s was chosen. The data recorded from GMRT has been converted to FITS and was analysed using Astronomical Image Processing System (AIPS). The details of observations/analysis is given elsewhere [6]. The 2 Yadav et al. Figure 1. The RXTE/ASM flux (left axis) and hardness ratio (right axis) during June 17 to July 3, 2001(top panel). The observed 1.28 GHz flux from GMRT and 15 GHz flux from Ryle telescope are shown in bottom panel (5min bin) for the same duration. radio observations at 15 GHz are from the Ryle telescope at Cambridge. Details of the instrument and analysis can be found elsewhere [9]. The RXTE/ASM flux and X-ray hardness ratio (5-12 keV/1.5-5 keV) are shown in the top panel of Figure 1. The radio lightcurves at 1.28 and 15 GHz produced at 5 min interval are shown in the bottom panel of Figure 1. GRS 1915+105 exhibited significant radio emission on all days, except on June 27, when the source showed much weaker radio emission (∼ 5 mJy). Figure 1 brings out following important points: 1. The radio fluxes at 1.28 and 15 GHz are consistent and suggest a flat radio spectrum during these observations except during the large radio flare on June 20-24. The flux and decay time of this flare suggest that it belongs to the class of relativistic jets. The higher flux at 1.28 GHz is consistent with steep spectrum during the decay of this flare. 2. The X-ray hardness ratio changed from ∼ 0.7 to ∼ 0.85 around June 30 prior to when source went to the plateau state on July 3. The RXTE/ASM flux becomes stable at higher hardness ratio. The X-ray hardness ratio suggests that the source remains in the low-hard state during this period. The radio light curve observed on June 28 is shown in Figure 2. The flux density was ∼ 5 mJy at about UTC 17 hour and reached gradually to 50 mJy at UTC 24 hour. When the observations resumed on 29 (not shown here), the source was "caught" at 70 mJy at UTC 16 hour, and the flux started decaying Workshop Proceedings 3 y J m 50 40 30 20 10 0 17 18 19 20 21 22 23 24 25 June 28 (hours) Figure 2. The radio light curve from GMRT at 1.28 GHz (mJy) observed on June 28, 2001 (5 min. bin). slowly to the value of 10 mJy at UTC 24 hour. This is probably the first detail observation of the precursor flares to the plateau state with rise and decay times of over six hours. It is also important to note that these broad flares and the change in the hardness ratio around June 30 set the stage for the plateau state. These broad flares consist of mini (baby) radio flares riding above rising/decaying radio emission. In contrast, the radio lightcurve observed on June 30 shows mini (baby) flares riding above almost smooth radio emission (Figure 3). y J m 35 30 25 20 15 10 5 0 15 16 17 18 20 21 19 June 30 (hours) 22 23 24 25 Figure 3. The radio light curve from GMRT at 1.28 GHz (mJy)observed on June 30, 2001 ( 5 min. bin). 4 Yadav et al. These isolated baby flares are modeled as adiabatically expanding synchrotron clouds ejected from the accretion disk [6]. One such flare along with model fit data is shown in Figure 4. This model provides estimate of the spectrum index from single frequency observations and successfully explains the observed delay times between different frequencies. The radio emission during these observations is consistent with flat radio spectrum which is in agreement with observed flux at 15 GHz. ) y J m ( x u l f z H G 3 . 1 35 30 25 20 15 10 5 0 20.1 20.2 20.3 20.4 20.5 20.6 Time on June 30 (hours) Figure 4. Isolated radio flare observed with GMRT at 1.28 GHz on 30 June, 2001 with model fit (for details see text). Acknowledgments We thank the staff of the GMRT that made these observations possible. GMRT is run by the National Center of Astrophysics of the Tata Institute of Fundamental Research. We also thank RXTE/ASM team for making their data public. References 494, L61. 1. Dhawan, V., Mirabel, I. F., Rodri´guez, L. F., 2000, ApJ, 543, 373. 2. Eikenberry, S. S., Matthews, K., Morgan, E. H., Remillard, R. A., Nelson, R. W.,1998, ApJ, 3. Fender, R. P., et al. 1999, MNRAS, 304, 865. 4. Mirabel, I. F., Rodri´guez, L. F., 1994, Nature, 371, 46. 5. Harmon, B. A., Deal, K. J., Paciesas, W. S., Zhang, S. N., Robinson, C. R., Gerad, E., Rodriguez, L. F., Mirabel, I. F., 1997, ApJ, 477, L85. Ishwara-Chandra, C. H., Yadav, J. S., Pramesh Rao, A., 2002, A&A, 388, L33. 6. 7. Mirabel, I. F., Dhawan, V., Chaty, S., Rodri´guez, L. F., Marti, J., Robinson, C. R., Swank, 8. Muno, M. P., Remillard, R. A., Morgan, E. H., Waltman, E. B., Dhawan, V., Hjellming, R. J., Geballe, T., 1998, A&A, 330, L9. M., Pooley, G. G., 2001, ApJ, 556, 515. 9. Pooley, G. G., Fender, R. P., 1997, MNRAS, 292, 925. 10. Swarup, G., Ananthakrishnan, S., Kapahi, V. K., Rao, A. P., Subrahmanya, C. R., Kulka- rni, V. K., 1991, Curr. Sci., 60, 95. 11. Yadav, J. S., 2001, ApJ, 548, 876.
astro-ph/0403352
2
0403
2004-03-16T20:29:22
Properties of Cold Dark Matter Halos at z>6
[ "astro-ph" ]
We compute the properties of dark matter halos with mass $10^{6.5}-10^9\msun$ at redshift $z=6-11$ in the standard cold dark matter cosmological model, utilizing a very high resolution N-body simulation. We find that dark matter halos in these mass and redshift ranges are significantly biased over matter with a bias factor in the range 2-6. The dark matter halo mass function displays a slope of $2.05\pm 0.15$ at the small mass end. We do not find a universal dark matter density profile. Instead, we find a significant dependence of the central density profile of dark matter halos on halo mass and epoch with $\alpha_0=0.4-1.0$; the high-mass ($M\ge 10^8\msun$) low-redshift ($z\sim 6$) halos occupy the high end of the range and low-mass ($M\sim 10^{7}\msun$) high-redshift ($z\sim 11$) halos occupy the low end. Additionally, for fixed mass and epoch there is a significant dispersion in $\alpha_0$ due to the stochastic assembly of halos. Our results fit a relationship of the form $\alpha_0=0.75((1+z)/7.0)^{-1.25}(M/10^7\msun)^{0.11(1+z)/7.0}$ with a dispersion about this fit of $\pm 0.5$ and no systematic dependence of variance correlated with environment. The median spin parameter of dark matter halos is $0.03-0.04$ but with a large lognormal dispersion of $\sim 0.4$. Various quantities are tabulated or fitted with empirical formulae.
astro-ph
astro-ph
Submitted to ApJ Properties of Cold Dark Matter Halos at z > 6 Renyue Cen1, Feng Dong2, Paul Bode3, and Jeremiah P. Ostriker4 ABSTRACT We compute the properties of dark matter halos with mass 106.5 − 109 M⊙ at redshift z = 6 − 11 in the standard cold dark matter cosmological model, utilizing a very high resolution N-body simulation. We find that dark matter halos in these mass and redshift ranges are significantly biased over matter with a bias factor in the range 2 − 6. The dark matter halo mass function displays a slope of 2.05±0.15 at the small mass end. We do not find a universal dark matter density profile. Instead, we find a significant dependence of the central density profile of dark matter halos on halo mass and epoch with α0 = 0.4 − 1.0; the high-mass (M ≥ 108 M⊙) low-redshift (z ∼ 6) halos occupy the high end of the range and low-mass (M ∼ 107 M⊙) high-redshift (z ∼ 11) halos occupy the low end. Additionally, for fixed mass and epoch there is a significant dispersion in α0 due to the stochastic assembly of halos. Our results fit a relationship of the form α0 = 0.75((1 + z)/7.0)−1.25(M/107 M⊙)0.11(1+z)/7.0 with a dispersion about this fit of ±0.5 and no systematic dependence of variance correlated with environment. The median spin parameter of dark matter halos is 0.03 − 0.04 but with a large lognormal dispersion of ∼ 0.4. Various quantities are tabulated or fitted with empirical formulae. Subject headings: ture of universe -- quasars: absorption lines cosmology: theory -- intergalactic medium -- large-scale struc- 1Princeton University Observatory, Princeton University, Princeton, NJ 08544; [email protected] 1Princeton University Observatory, Princeton University, Princeton, NJ 08544; [email protected] 1Princeton University Observatory, Princeton University, Princeton, NJ 08544; [email protected] 1Princeton University Observatory, Princeton University, Princeton, NJ 08544; [email protected]; Institute for Astronomy, Cambridge University, Cambridge, England; [email protected] -- 2 -- 1. Introduction The reionization epoch is now within the direct observational reach thanks to rapid recent observational advances in two fronts -- optical quasar absorption from Sloan Digi- tal Sky Survey (SDSS) (Fan et al. 2001; Becker et al. 2001) and the Wilkinson Microwave Anisotropy Probe (WMAP) experiment (Kogut et al. 2003). The picture painted by the combined observations, perhaps not too surprisingly, strongly suggests a complex cosmologi- cal reionization process, consistent with the double reionization scenario (Cen 2003). It may be that this is the beginning of a paradigm shift in our focus on the high redshift universe: the star formation history of the early universe can now be observationally constrained. It thus becomes urgent to theoretically explore galaxy and star formation process at high redshift in the dark age (z ≥ 6). In the context of the standard cold dark matter model it is expected that stars within halos of mass 107 − 109 M⊙ at high redshift play an important, if not dominant, role in determining how and when the universe was reionized. Furthermore, these fossil halos may be seen in the local universe as satellites of giant galaxies. This linkage may potentially provide a great leverage to nail down the properties of the high redshift galaxies. In this paper, as a step towards understanding galaxy formation at high redshift, we investigate the properties of dark matter halos at z ≥ 6, using very high resolution TPM N-body (Bode et al. 2001; Bode & Ostriker 2003) simulations. While there is an extensive literature on properties of halos at low redshift, there is virtually no systematic study of dark halos at z ≥ 6. The LCDM simulation has a comoving box size of 4h−1Mpc with 5123 = 108.2 particles, a particle mass of mp = 3.6× 104 h−1 M⊙, and comoving gravitational softening length of 0.14 h−1kpc. These resolutions allow us to accurately characterize the properties of halos down to a mass 106.5 h−1 M⊙ (having about 100 particles within the virial radius). The outline of this paper is as follows. The simulation details are given in §2. In §3 we quantify properties of dark matter halos in the mass range 106.5 − 109 M⊙, including the mass function, bias and clustering properties, density profile distribution, angular momentum spin parameter distribution, internal angular momentum distribution and peculiar velocity distribution. We conclude in §4. 2. The Simulation A standard spatially flat LCDM cosmology was chosen, with Ωm = 0.27 and ΩΛ = 0.73; the Hubble constant was taken to be 70 km/s/Mpc. The initial conditions were created using the GRAFIC2 package by Bertschinger (2001). The matter transfer function was calculated -- 3 -- with the included Boltzmann integrator (Ma & Bertschinger 1995), using Ωbh2 = 0.211 for the baryon fraction and σ8 = 0.73 for the normalization of the matter power spectrum. The simulation contained N = 5123 particles in a comoving periodic box 4h−1Mpc on a side, making the particle mass mp = 3.57 × 104 h−1M⊙. The starting redshift was z = 53, and the system was evolved down to z = 6. The evolution was carried out with the parallel Tree -- Particle -- Mesh code TPM (Xu 1995; Bode, Ostriker, & Xu 2000; Bode & Ostriker 2003), using a 10243 mesh. The evolution took 1150 PM steps, with particles in dense regions taking up to 19,500 steps. The run was carried out using up to 256 processors on the Terascale Computing System at Pittsburgh Supercomputing Center. The spline softening length in the tree portion of the code was set to ǫ = 0.14 h−1kpc. With this softening length, relaxation by z = 6 inside the core of a collapsed halo (assuming an NFW density distribution with c = 12) will not be significant over the course of the simulation for those objects containing more than 100 particles. The opening angle in the Barnes-Hut criterion used by TPM was θ = 0.577, and the time step parameter η = 0.3; also, the initial values for locating trees were A = 2.0 and B = 12.5 -- see Bode & Ostriker (2003) for details. In the TPM code, not all regions are treated at full resolution. The limiting density (above which all cells are put into trees for increased resolution) rises with time. By the end of this run, all cells containing more than 18 particles are still being followed at full resolution. Thus this factor is not important if the analysis is limited to halos with over 100 particles. Dark matter halos are identified using DENMAX scheme (Bertschinger & Gelb 1991), smoothing the density field with a Gaussian length of 300h−1kpc. In computing all quantities we include all particles located inside the virial radius of a halo. 3. Results 3.1. Pictures First, we present visually a distribution of the dark matter mass and dark matter halos of varying masses. Figure 1 shows the distribution of dark matter particles projected onto the x-y plane. Figures (2a,b,c,d,e) show the distributions of dark matter halos with masses greater than (106, 106.5, 107, 107.5, 108) M⊙, respectively, at z = 6. The progressively stronger clustering of more massive halos is clearly visible in the display but we will return to the clustering properties more quantitatively in §3.3. It is also noted that voids are progressively more visible in the higher mass halos than in low mass halos. -- 4 -- Fig. 1. -- The distribution of dark matter particles at z = 6 projected onto the x-y plane (0.25% of the total). -- 5 -- M > 106 h−1 M⊙ 2. -- The distributions of all dark matter halos with masses greater Fig. (106, 106.5, 107, 107.5, 108) M⊙, respectively, at z = 6. than -- 6 -- M > 106.5 h−1 M⊙ Fig. 2. -- Continued. -- 7 -- M > 106 h−1 M⊙ Fig. 2. -- Continued. -- 8 -- M > 107.5 h−1 M⊙ Fig. 2. -- Continued. -- 9 -- M > 108 h−1 M⊙ Fig. 2. -- Continued. -- 10 -- 3.2. Dark Matter Halo Mass Function Figure 3 shows the halo mass functions at four redshifts. Table 1 summarizes the fitting parameters for a Schechter function of the form n(M)dM = n0( M M∗ )−α exp (−M/M∗) dM M∗ . (1) We see that the Schechter function provides a good fit to the computed halo mass function. The faint end slope is α = 2.05 ± 0.15, consistent with the expectation from Press-Schechter theory (Press & Schechter 1974). While there appears to be a slight steepening of the slope at the low mass end from −1.9 to −2.2 from z = 6 to z = 11.1, it is unclear, however, how significant this trend is, given the adopted, somewhat degenerate fitting formula. The turnover at Mh ∼ 106.5 M⊙ indicates the loss of validity of our simulation at the low mass end. 3.3. Bias of Dark Matter Halos We characterize the relative distribution of halos over the total dark matter distribution by the following relation: nh < nh > = b(M, z)( ρm < ρm > )c(M,z), (2) where nh and < nh > are the halo density and mean halo density; ρm and < ρm > are ρm the mass density and mean mass density; c(M, z) is fixed to be unity at <ρm> > 1. This empirical fitting formula is motivated by the found result that there appears to be a break in nh <nh> at <ρm> ∼ 1. Tables 2 and 3 list the parameters b(M, z) and c(M, z). The smoothing length used here is 0.3 h−1 Mpc. Figure 4 shows four typical cases to indicate the goodness <ρm> )3−7, a of the fitting formula. At ρm <ρm> < 1 our results (Table 2) indicate ρm nh <nh> ∝ ( ρm Table 1. Halo Mass Function Fitting Parameters Parameters z=6.0 z=7.4 z=9.08 z=11.096 n0 (h3 Mpc−3) α M∗ (h−1 Mpc) 0.85 1.9 8 × 108 1.20 2.0 6 × 108 1.75 2.1 4 × 108 2.40 2.2 2 × 108 -- 11 -- Fig. 3. -- The halo mass functions at four redshifts. Dotted lines represent the fitted Schechter functions with parameters summarized in Table 1. -- 12 -- 10 1 0.1 0.1 1 10 0.1 1 10 10 1 0.1 10 1 0.1 10 1 0.1 0.1 1 10 0.1 1 10 Fig. 4. -- Bias of dark matter halos in four randomly selected cases upon different mass and redshift. The darkened lines show the fitting formulae of the bias relation. -- 13 -- Table 2. Halo Bias : b(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 >106.0 >106.5 >107.0 >107.5 >108.0 1.16±0.12 1.21±0.13 1.37±0.19 1.72±0.32 2.08±0.55 1.36±0.18 1.45±0.17 1.60±0.23 2.12±0.47 2.58±0.90 1.70±0.20 2.09±0.30 2.77±0.40 3.46±0.87 4.43±1.47 2.46±0.23 3.05±0.33 3.92±0.90 6.17±3.17 9.90±4.58 Table 3. Halo Bias : c(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 >106.0 >106.5 >107.0 >107.5 >108.0 3.00±0.55 3.27±0.71 3.43±0.70 3.93±1.70 4.18±2.28 3.80±0.79 4.20±1.04 4.37±1.07 4.98±2.96 5.61±4.68 4.14±0.80 4.41±1.31 4.97±2.59 5.09±3.31 6.32±6.12 4.90±1.49 5.52±2.43 7.33±6.29 7.41±9.09 6.77±8.90 -- 14 -- Fig. 5. -- shows the ratio of dark matter halo correlation function over the dark matter mass correlation function at four redshifts, z = 6.0 (top right panel), z = 7.4 (top left), z = 9.1 (bottom right) and z = 11 (bottom left). In each panel six curves are shown for halos more massive than 106 M⊙ 106.5 M⊙ 107 M⊙ 107.5 M⊙ 108 M⊙ and 108.5 M⊙, respectively. -- 15 -- Fig. 6. -- shows the bias as a function of halo mass at redshifts z = (6, 7.4, 9.1, 11.1) at separation of 1h−1Mpc as symbols (x). The curves are computed using the analytic method of Mo & White (1996). -- 16 -- rather rapid drop. As expected, the drop-off is more dramatic for larger halos, as visible in Figure 2. This implies that at z > 6 halos are unlikely to be found in underdense regions (on a scale of ∼ 0.3Mpc/h). The increase of c(M, z) with redshift implies that voids are emptier at higher redshifts. Another way to characterize the relative distribution of dark matter halos over mass is to compute the ratio of the correlation functions, which are shown in Figure 5. The correlation function ξ(r) is calculated by counting the number of pairs of either particles or halos at separation r (using logarithmically spaced bins) and comparing that number to a Poisson distribution. It can be seen that the bias falls in the range 2 − 6, with the trend that the more massive halos are more biased and at a fixed mass halos at higher redshifts are more biased, as expected. Figure 6 recollects the information in Figure 5 and shows the bias as a function of halo mass at four different redshifts at the scale chosen to be 1h−1Mpc. The agreement between our computed results and that using analytic method of Mo & White (1996) is good, indicating that the latter is valid for objects at scales and redshifts of concern here. 3.4. Dark Matter Halo Density Profile We use a variant fitting formula based on the NFW (Navarro, Frenk, & White 1997) profile: ρr = ρs ( r r−2 )α(1 + r r−2 )4−2α . (3) Note that for NFW profile, α = 1. An important difference, however, is the scaling radius used. We use the radius where the logarithmic slope of the density profile is −2, r−2, instead of the more conventional "core" radius. This is a two parameter fitting formula, α and r−2, while ρs is a function of α and r−2 at a fixed redshift, since the overdensity interior to the virial radius rv is assumed to be known. This fitting formula is intended for the range in radius r ≤ rv only. We fit the density profile of each halo using the least squares method. Four randomly selected examples of such profiles along with the fitted curve using Equation (3) are shown in Figure 7, indicating reasonable fits in all cases. Both fitting parameters, α and r−2, however, display broad distributions. We found that there is only weak correlation between α and r−2. Figure 8 shows histograms for the distributions of α, for four typical cases. We fit α -- 17 -- 0.1 1 0.1 1 0.1 1 0.1 1 Fig. 7. -- Density profiles of four randomly selected halos over a mass range at z = 6. The dotted lines represent the best-fit modified NFW relation given in Section 3. -- 18 -- Fig. 8. -- The distributions of inner slope parameter α in the halo density profile fitting for four randomly selected cases upon different mass and redshift. The Gaussian fits are shown as smooth curves. -- 19 -- distributions using a Gaussian distribution function: P (α) = 1 √2πσα exp (− (α − α0)2 2σα 2 ). (4) The Gaussian fits are shown as smooth curves in Figure 8, demonstrating that the proposed Gaussian fits are good. Tables 4,5 list fitting parameters α0 and σα0, respectively. We recollect the data in Table 4 and shows in Figure 9 the median inner density slope as a function of dark matter halo different mass at four different redshifts (symbols). The curves in Figure 9 are empirical fits using the following formula α0 = 0.75((1 + z)/7.0)−1.25(M/107 M⊙)0.11(1+z)/7.0. (5) It is seen that this fitting formula provides a reasonable fit for the simulated halos. Figure 10 shows histograms for the distributions of r−2 for four typical cases. We fit r−2 distributions using a lognormal function: P (r−2) = 1 r−2√2πσr−2 exp (− (ln r−2 − ln r0 −2)2 2σr−2 ), (6) which are seen to provide reasonable fits to the data in Figure 10. Tables 6,7 list fitting parameters r0 −2 and σr−2, respectively. We find no visible correlation between α and r−2, as shown in Figure 11. An important point to note is that the average slope of the density profiles of small halos ranges from 1.0 to 0.4 from z = 6 to z = 11 (Table 4 and Figure 8), which is somewhat shallower than the universal density profile found by Navarro et al. (1997). Our results at z = 9 − 11 are in good agreement with Ricotti (2003), who simulated a 1h−1Mpc box for small halos at z = 10. If the theoretical argument for the dependence of halo density profile on the slope of the initial density fluctuation power spectrum (Syer & White 1988; Subramanian, Cen, & Ostriker 2000) is correct, as adopted by Ricotti (2003) to explain the Table 4. Density Profile : α0(M, z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.0 - 107.5 107.5 - 108.0 > 108.0 0.81±0.01 0.90±0.01 1.09±0.02 0.63±0.01 0.75±0.02 0.91±0.02 0.50±0.01 0.62±0.02 0.77±0.03 0.42±0.01 0.50±0.04 0.65±0.06 -- 20 -- 11 0.9 0.8 0.7 0.6 0.5 0.4 Fig. 9. -- shows the median inner density slope as a function of dark matter halo different mass at four different redshifts (symbols). The curves are fits (Equation 5). Table 5. Density Profile : σα(M, z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.0 - 107.5 107.5 - 108.0 > 108.0 0.58±0.01 0.43±0.01 0.41±0.01 0.54±0.01 0.45±0.01 0.41±0.02 0.52±0.01 0.46±0.02 0.41±0.03 0.51±0.01 0.56±0.03 0.38±0.06 -- 21 -- Fig. 10. -- The distributions of radius parameter r−2 in the halo density profile fitting for four randomly selected cases upon different mass and redshift. The lognormal fits are shown as smooth curves. -- 22 -- dependence of inner density slope on halo mass, it then follows that the neglect of density fluctuations on scales larger than his box size of 1h−1Mpc in Ricotti (2003) would make the density profiles of the halos in his simulation somewhat shallower than they should be in the CDM model. Thus, the difference in the box size (1h−1Mpc in Ricotti 2003 versus 4h−1Mpc for our simulation box) would have expected to result in a slightly steeper inner slope in our simulation, which is indeed the case. Another point to note, which is not new but not widely known, is that there is a large dispersion in the inner slope of order 0.5 due to the intrinsically stochastic nature of halo assembly. This was found earlier by Subramanian et al. (2000). Therefore, while a "universal" profile is informative in characterizing the mode, a dispersion would be needed to give a full account. This is particularly important for applications where the dependence on the inner slope is very strong, e.g., strong gravitational lensing. More relevant for our case of small halos is that, for example, the fraction of small halos with inner slope close to zero (i.e., flat core) is non-negligible at z = 6. A proper statistical comparison with observations of local dwarf galaxies, however, is not possible with the current simulation without evolving small galaxies to z = 0. Also worthwhile is to understand whether or not there is some dependence of the inner slope of the density profile on the central density of a halo or the environmental density where a halo sits. In Figures 12 and 13 we show the correlation between the inner slope of the density profile, α, and the central density of the halo, and between α and the environmental density, respectively. We find no visible correlations between either pair of quantities. But a relationship between halo shape and environment might have been missed due to stochastic variations. Thus, we have checked to see if deviation ∆α, between computed and predicted (based on mass and epoch using equation 5) slope exists. Figure 14 shows this and no correlation is seen. We note that the abundances of halos in the mass range of interest here are on the rise in the redshift range considered (Lacey & Cole 1993), as evident in Figure 3. During this period halos considered the inner density profiles of halos steepen with Table 6. Density Profile : r0 −2(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.0 - 107.5 107.5 - 108.0 > 108.0 0.35±0.01 0.35±0.01 0.34±0.01 0.40±0.01 0.38±0.01 0.35±0.01 0.43±0.01 0.39±0.01 0.34±0.01 0.44±0.01 0.37±0.01 0.34±0.02 -- 23 -- Fig. 11. -- A scatter plot of the inner slope of the density profile, α, versus r−2 for all the halos with M > 106.5 h−1M⊙ at z = 6. -- 24 -- Fig. 12. -- A scatter plot of the inner slope of the density profile, α, versus the halo central density, defined as the density at r < 0.2rv. for all the halos with M > 106 h−1M⊙ at z = 6. -- 25 -- Fig. 13. -- A scatter plot of the inner slope of the density profile, α, versus the environmen- tal density, defined as the dark matter density smoothed by a gaussian window of radius 0.3h−1Mpc, for all the halos with M > 106 h−1M⊙ at z = 6. -- 26 -- time, consistent with the increase of logarithmic slope of the power spectrum with time, that corresponds to the evolving nonlinear mass scale. However, at some lower redshift not probed here, low mass halos will cease to form. Subsequently, the evolution of halo density profile may show distinct features and some conceivable correlations, not seen in Figures 12 and 13, may show up. We will study this issue separately. Since the central density of a halo may be considered a good proxy for the formation redshift of the central region, the non-correlation between α and the central density indicates that, for halos of question here, the subsequent process of accretion of mass onto halos is largely a random process, independent of the density of the initial central "seed". The fact that halos of different masses show a comparable range of central density (not shown in figures) suggest that halos of varying masses form nearly simultaneously, dictated by the nature of the cold dark matter power spectrum at the high-k end; i.e., density fluctuations on those scales involved here depend weakly (logarithmically) on the mass. The non-correlation between the inner slope of the halo density profile and the environmental density may be interpreted in the following way. One may regard regions of different overdensities as local mini-universes of varying density parameters. The independence of the inner slope on local density is thus consistent with published results that the halo density profiles in universes of different ΩM do not significantly vary. 3.5. Spin Parameter For Dark Matter Halos We compute the spin parameter defined as JE1/2 GM 5/2 λ ≡ (7) (Peebles 1969), where G is the gravitational constant; M is the total mass of the dark matter halo; J is the total angular momentum of the dark matter halo; E is the total energy of Table 7. Density Profile : σr−2(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.0 - 107.5 107.5 - 108.0 > 108.0 0.34±0.01 0.31±0.01 0.33±0.01 0.32±0.01 0.31±0.01 0.25±0.01 0.29±0.01 0.27±0.01 0.26±0.02 0.27±0.01 0.27±0.01 0.27±0.05 -- 27 -- Fig. 14. -- A scatter plot of the difference between measured inner slope of the density profile, α, and the fitted inner slope using equation (5), versus the environmental density. -- 28 -- Fig. 15. -- The distributions of halo spin parameter λ for four randomly selected cases upon different mass and redshift. The modified lognormal fits are shown as smooth curves. -- 29 -- Fig. 16. -- A scatter plot of the spin parameter, λ, versus the halo central density, defined as the density at r < 0.2rv. for all the halos with M > 106 h−1M⊙ at z = 6. -- 30 -- Fig. 17. -- A scatter plot of the spin parameter, λ, versus the environmental density, defined as the dark matter density smoothed by a gaussian window of radius 0.3h−1Mpc, for all the halos with M > 106 h−1M⊙ at z = 6. -- 31 -- Fig. 18. -- Comparison of the halo density profile slope parameter α versus spin parameter λ. No correlation between the two is observed. -- 32 -- Table 8. Spin Parameter : λmedian(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.00 - 107.25 107.25 - 107.50 107.50 - 108.00 > 108.00 0.043±0.001 0.042±0.001 0.041±0.001 0.035±0.001 0.043±0.001 0.041±0.001 0.039±0.001 0.033±0.001 0.042±0.001 0.037±0.001 0.035±0.001 0.031±0.002 0.040±0.001 0.037±0.001 0.035±0.002 0.031±0.007 Table 9. Spin Parameter : σλ(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.00 - 107.25 107.25 - 107.50 107.50 - 108.00 > 108.00 0.43±0.01 0.43±0.01 0.42±0.01 0.40±0.01 0.44±0.01 0.42±0.01 0.41±0.01 0.39±0.02 0.44±0.01 0.38±0.01 0.41±0.01 0.31±0.04 0.41±0.01 0.41±0.02 0.32±0.03 0.43±0.15 Table 10. Angular Momentum Profile : A0(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.0 - 107.5 107.5 - 108.0 > 108.0 0.51±0.03 0.55±0.06 0.56±0.09 0.49±0.04 0.52±0.12 0.51±0.21 0.44±0.06 0.48±0.14 0.49±0.24 0.44±0.08 0.50±0.16 0.45±0.32 Table 11. Angular Momentum Profile : σA(M,z) Halo Mass (h−1 M⊙) z=6.0 z=7.4 z=9.08 z=11.096 107.0 - 107.5 107.5 - 108.0 > 108.0 0.27±0.03 0.26±0.05 0.27±0.08 0.28±0.04 0.29±0.10 0.35±0.18 0.27±0.06 0.29±0.11 0.29±0.16 0.25±0.07 0.26±0.12 0.36±0.23 -- 33 -- the dark matter halo; all quantities are computed within the virial radius. We fit the λ distributions using a modified lognormal function: P (λ) = 1 (λ + ǫ0)√2πσλ exp (− [ln(λ + ǫ0) − ln λ0]2 2σλ ), (8) where ǫ0 is fixed to be 0.0125, which is determined through experimentation. The modified lognormal fits for λ are shown as smooth curves in Figure 15; the goodness of the fits is typical. Tables 8,9 lists fitting parameters λ0 and σλ, respectively. Note that the median value of λ is λmed = λ0 − ǫ0. We see that the typical spin parameter has a value 0.03−0.04. However, the distribution of the spin parameter among halos is very broad, with a lognormal dispersion of ∼ 0.4. This implies that consequences that depend on the spin of a halo are likely to be widely distributed even at a fixed dark matter halo mass. Such consequences may include the size of a galactic disk and correlations between dark matter halo spin (conceivable misalignment between spin of gas and spin of dark matter would complicate the situation) and other quantities. In Figures 16 and 17 we show the correlation between λ and the central density of the halo, and between λ and the environmental density, respectively. We find that there may possibly exist a weak correlations between λ and the central density of the halo, in the sense that halos with higher central densities (or equivalently earlier formation times) have lower λ, with a very large scatter, whereas no correlation is discernible between λ and the environmental density. Finally, Figure 18 show the relation between λ and α, where no correlation is visible. 3.6. Angular Momentum Profile For Dark Matter Halos Next, we compute the angular momentum profiles for individual dark matter halos. We then fit the each angular momentum profile by the following function in the small j regime: M(< j) Mv = A j j0 + M(< 0) Mv , (9) where A and M(< 0) are two fitting parameters; Mv is the virial mass and j0 ≡ J/Mv = λGM 3/2/E1/2. In order to compute the angular momentum profile an appropriate smoothing window needs to be applied to dark matter particles. We find that M(< 0)/Mv varies for different smoothing scales, with typical values around 0.2, while A remains roughly constant for each individual halo. However, A is broadly distributed for all dark matter halos. Figure 19 -- 34 -- Fig. 19. -- The distributions of slope parameter A in the halo angular momentum profile fitting for four randomly selected cases upon different mass and redshift. The lognormal fits are shown as smooth curves. -- 35 -- Fig. 20. -- The distributions of dark matter halo peculiar velocity for four mass bins at z = 6. The lognormal fits are shown as smooth curves. -- 36 -- show histograms for the distributions of A0, for four randomly selected cases. We fit the A distribution using a modified lognormal function: P (A) = 1 (A + ǫA)√2πσA exp (− [ln(A + ǫA) − ln A0]2 2σA ), (10) where ǫA = 0.4 is fixed through experimentation. Tables 10,11 list fitting parameters A0 and σA, respectively. In general we find our fitting formula (Equation 10) provides a good fit for each individ- ual halo. The distribution of matter at small j is most relevant for the formation of central objects, such as black holes (e.g., Colgate et al. 2003) or bulges (e.g., D'Onghia & Burkert 2004). Our calculation indicates that the fraction of mass in a halo having specific angular momentum less than a certain value is roughly 0.5 times the ratio of that value over the average specific angular momentum of the halo. 3.7. Bulk Velocity of Dark Matter Halos Finally, we compute the peculiar velocity of dark matter halos. We find that the distribu- tion once again can be fitted by lognormal distributions as equation (9). In order to provide a good fit for the results at z = 6 shown in Figure 20, it is found that ǫ = 40km/s in equation (8) with median velocity vm ∼ 38 ± 2km/s and lognormal dispersion σv = 0.22 ± 0.01. We caution, however, the absolute value of the peculiar velocity of each halo, unlike the quan- tities examined in previous subsections, may be significantly affected by large waves not present in our simulation box. Adding missing large waves should increase the zero point ǫ to a larger value. Therefore, the peculiar velocity shown should be treated as a lower limit. In other words, expected peculiar velocity of dark matter halos at these redshifts are likely in excess of 30 − 40km/s. 4. Conclusions Using a high resolution TPM N-body simulation of the standard cold dark matter cosmological model with a particle mass of mp = 3.57 × 104 h−1M⊙ and a softening length of ǫ = 0.14 h−1kpc in a 4h−1Mpc box, we compute various properties of dark matter halos with mass 106.5 − 109 M⊙ at redshift z = 6 − 11. We find the following results. (1) Dark matter halos at such small mass at high redshifts are already significantly biased over matter with a bias factor in the range 2 − 6. -- 37 -- (2) The dark matter halo mass function displays a slope at the small end 2.05 ± 0.15. (3) The central density profile of dark matter halos are found to be in the range (0.4 − 1.0) well fitted by α0 = 0.75((1 + z)/7.0)−1.25(M/107 M⊙)0.11(1+z)/7.0 with a dispersion of ±0.5, in rough agreement with the theoretical arguments given in Ricotti (2003) and Subramanian et al. (2000). (4) The median spin parameter of the dark matter halos is 0.03− 0.04 but with a lognormal dispersion of ∼ 0.4. The angular momentum profile at the small end is approximately linear with the fraction of mass in a halo having specific angular momentum less than a certain value is roughly 0.5 times the ratio of that value over the average specific angular momentum of the halo. (5) The dark matter halos move at a typical velocity in excess of 30 − 40km/s. This research was supported in part by AST-0206299 and NAG5-13381. The computa- tions were performed on the National Science Foundation Terascale Computing System at the Pittsburgh Supercomputing Center. REFERENCES Becker, R.H., et al. 2001, AJ, 122, 2850 Bertschinger, E.,& Gelb, J.M. 1991, Computers in Physics, 5, 164 Bertschinger, E. 2001, ApJS, 137, 1 Bode, P., & Ostriker, J.P. 2003, ApJS, 145, 1 Bode, P., Ostriker, J.P., & Xu, G. 2000, ApJS, 128, 561 Cen, R. 2003, ApJ, 591, 12 Colgate, S.A., Cen, R., Li, H., Currier, N., & Warren, M.S. 2003, ApJL in press, astro- ph/0310776 D'Onghia, E., & Burkert, A. 2004, astro-ph/0402504 Fan, X., et al. 2001, AJ, 122, 2833 Lacey, C., & Cole, S. 1993, MNRAS, 262, 627. Ma, C.-P., & Bertschinger, E. 1995, ApJ, 455, 7 Mo, H.J., & White, S.D.M. 1996, MNRAS, 282, 347 Navarro, J., Frenk, C.S., & White, S.D.M. 1997, ApJ, 490, 493 -- 38 -- Peebles, P.J.E. 1969, ApJ, 155, 393 Press, W.H., & Schechter, P. 1974, ApJ, 187, 425 Peebles, P.J.E. 1993, Principles of Physical Cosmology (Princeton: Princeton University Press) Ricotti, M. 2003, MNRAS, 344, 1237 Syer, D., & White, S.D.M. 1998, MNRAS, 293, 337 Subramanian, K., Cen, R., & Ostriker, J.P. 2000, ApJ, 538, 528 Xu, G. 1995, ApJS, 98, 355 This preprint was prepared with the AAS LATEX macros v5.2.
astro-ph/0507338
1
0507
2005-07-14T14:44:57
Baryonic acoustic oscillations in simulated galaxy redshift surveys
[ "astro-ph" ]
Baryonic acoustic oscillations imprinted in the galaxy power spectrum provide a promising tool for probing the cosmological distance scale and dark energy. We present results from a suite of cosmological N-body simulations aimed at investigating possible systematic errors in the recovery of cosmological distances. We show the robustness of baryonic peaks against nonlinearity, redshift distortions, and mild biases within the linear and quasilinear region at various redshifts. While mildly biased tracers follow the matter power spectrum well, redshift distortions do partially obscure baryonic features in redshift space compared to real space. We calculate the statistical constraints on cosmological distortions from N-body results and compare these to the analytic results from a Fisher matrix formalism. We conclude that the angular diameter distance will be constrained as well as our previous Fisher matrix calculations while the Hubble parameter will be less constrained because of nonlinear redshift distortions.
astro-ph
astro-ph
Accepted by The Astrophysical Journal 7-11-2005 Preprint typeset using LATEX style emulateapj v. 26/01/00 5 0 0 2 l u J 4 1 1 v 8 3 3 7 0 5 0 / h p - o r t s a : v i X r a BARYONIC ACOUSTIC OSCILLATIONS IN SIMULATED GALAXY REDSHIFT SURVEYS Hee-Jong Seo, Daniel J. Eisenstein August 28, 2018 Steward Obsevatory, University of Arizona, 933 North Cherry Avenue, Tucson, AZ 85721 Accepted by The Astrophysical Journal 7-11-2005 ABSTRACT Baryonic acoustic oscillations imprinted in the galaxy power spectrum provide a promising tool for probing the cosmological distance scale and dark energy. We present results from a suite of cosmological N-body simulations aimed at investigating possible systematic errors in the recovery of cosmological distances. We show the robustness of baryonic peaks against nonlinearity, redshift distortions, and mild biases within the linear and quasilinear region at various redshifts. While mildly biased tracers follow the matter power spectrum well, redshift distortions do partially obscure baryonic features in redshift space compared to real space. We calculate the statistical constraints on cosmological distortions from N-body results and compare these to the analytic results from a Fisher matrix formalism. We conclude that the angular diameter distance will be constrained as well as our previous Fisher matrix calculations while the Hubble parameter will be less constrained because of nonlinear redshift distortions. Subject headings: cosmological parameters -- large-scale structure of universe -- cosmology: theory -- distance scale -- methods: N-body simulations 1. INTRODUCTION Baryons create a distinct oscillatory signature in the power spectrum of the large-scale structure of the uni- verse (Peebles & Yu 1970; Bond & Efstathiou 1984; Holtz- man 1989; Hu & Sugiyama 1996; Eisenstein & Hu 1998a). These baryonic acoustic oscillations have been seen in the anisotropies of cosmic microwave background (Miller et al. 1999; de Bernardis et al. 2000; Hanany et al. 2000; Halverson et al. 2002; Lee et al. 2001; Netterfield et al. 2002; Benoit et al. 2003; Bennett et al. 2003; Pearson et al. 2003) and recently in large galaxy redshift surveys (Eisen- stein et al. 2005; Cole et al. 2005). Because the density of baryons is less than that of cold dark matter, the oscil- lations in the matter power spectrum are weaker in am- plitude than those in the cosmic microwave background (hereafter CMB). However, whether in the CMB or in late- time structure, the oscillations define a constant comoving length scale in linear perturbation theory. In our previous paper (Seo & Eisenstein 2003, hereater SE03), we demonstrated that large galaxy redshift surveys can constrain the Hubble parameter and angular diame- ter distance to a precision of a few percent using the im- printed baryonic acoustic oscillations as a standard ruler. The physical scale of the oscillations can be determined from matter density and baryon density of the universe, which in turn are deduced from the shape and relative amplitude of the baryonic peaks in CMB anisotropy data (Eisenstein et al. 1998; Eisenstein 2003). One can then compare the physical scale with the observed length scales of oscillations in transverse and line-of-sight directions in galaxy redshift surveys to yield the angular diameter dis- tance and Hubble parameter at the given redshift. This in turn measures the evolution of dark energy as well as the spatial curvature and Ωm. See Blake & Glazebrook (2003), Linder (2003), Hu & Haiman (2003), Amendola et al. (2005), Cooray (2004), Dolney et al. (2004), and Mat- subara (2004) for similar studies. Because of its weak amplitude modulations in matter distribution, the baryonic features are susceptible to the erasure by the nonlinear coupling of Fourier modes (Jain & Bertschinger 1994; Meiksin, White, & Peacock 1999; Meiksin & White 1999; Scoccimarro et al. 1999), such as can result from nonlinear growth, nonlinear bias (Kaiser 1987; Coles 1993; Scherrer & Weinberg 1998; Dekel & La- hav 1999; Coles et al. 1999; Seljak 2000; White 2001) or nonlinear redshift distortions (Hamilton 1998; Hatton & Cole 1998; Seljak 2001; White 2001; Scoccimarro 2004). Nonlinear mode-coupling also produces additional small- scale power, thereby altering the shape of the power spec- trum and potentially disguising the locations of baryonic features. It is therefore crucial to examine whether the os- cillatory features are robust against various nonlinearities and whether the remaining ripples can effectively distin- guish a dilation in distance from other non-cosmological effects in the clustering of galaxies. In this paper, we present an N-body study of the effect of nonlinear growth, nonlinear redshift distortions, and halo bias on the detectability of acoustic oscillations. Pre- viously, Meiksin, White, & Peacock (1999) used N-body simulations to show the effects of nonlinearity on baryonic signatures in the present-day large-scale structure. They studied the effect of bias and redshift distortions for var- ious cosmological models. Our study extends their work to higher redshifts. We adopt a ΛCDM model consistent with Wilkinson Microwave Anisotropy Probe (WMAP) data (Spergel et al. 2003) and generate density fields at redshifts of 3, 1 and 0.3. We then investigate and quantify the erasure of baryonic features in the matter power spec- trum and biased galaxy power spectrum in real space and redshift space at those redshifts. We also attempt to re- move the nonlinear alteration from the shape of the power spectrum and thereby recover the underlying contrast of baryonic features. As this work was being completed, re- lated studies by Springel et al. (2005) and Angulo et al. (2005) appeared on this topic. In SE03, we used the Fisher information matrix to cal- culate predictions for the statistical constraints on dark 1 2 energy. We used a conservative choice of nonlinear scale kmax (=π/2R) by requiring σR ∼ 0.5 and excluded power in smaller scales from our analysis. For larger scales, we adopted a linear growth function, linear redshift distor- tions, and a linear bias model with an additive offset. In reality, as the transition from linear to nonlinear scales is not discrete, the effects of nonlinearity, i.e., mode coupling and more complicated scale-dependence, may mildly con- taminate the power spectrum even on large scales. Here, we use our N-body simulations to assess the impacts of these nonlinear effects on the baryonic features on large scales relative to the statistical constraints we calculated in SE03. We compare our N-body results with the choices of nonlinear scale in SE03. Our results will provide further guidance for linear approximations in various studies with baryonic physics. Finally we want to investigate the distance constraints available from galaxy surveys taking account of the full N- body effect. The nonlinear effects not only inhibit us from detecting the weak baryonic signatures on small scales but also increase the statistical variance of power spectrum above the Gaussian estimates. We perform χ2 analysis on our simulated power spectra to fit to the cosmological dis- tances and compare the constraints with those in SE03. The result will show whether we can deduce the informa- tion on cosmological distances from the power spectrum altered by the nonlinear effect. In § 2 we describe the parameters of our cosmological N-body simulations. In § 3 we present the effect of nonlin- ear growth and redshift distortions on baryonic features in the matter power spectrum. In § 3.3 we remove the effect of nonlinearity on the broadband shape and study the re- sulting features. In § 4 we study the baryonic signatures in biased power spectra in real space and redshift space. In § 5 we present the errors on cosmological distances re- sulting from χ2 analysis. 2. COSMOLOGICAL N-BODY SIMULATIONS We run a series of cosmological N-body simulations us- ing the Hydra code (Couchman, Thomas, & Pearce 1995) in collisionless P3M mode. We use the CMBfast (Seljak & Zaldarriaga 1996; Zaldarriaga et al. 1998; Zaldarriaga & Seljak 2000) linear power spectrum to generate many ini- tial Gaussian random density fields at redshift of 49 and evolve them to redshifts of 3, 1, and 0.3. The cosmological parameters we use to generate the initial fields are Ωm = 0.27, ΩX = 0.73, Ωb = 0.046, h = 0.72, and n = 0.99. The initial fields are normalized by requiring σ8 = 0.9 at z = 0 and assuming a linear growth function. Each simulation box represents Vbox = 5123h−3 Mpc3 and contains 2563 dark matter particles (∼ 8.28 × 1011Msun/particle). We compute gravity using 2563 force grids with a Plummer softening length of 0.2 h−1 Mpc. We use 51 simulations at z = 1 and 0.3, and 30 simulations at z = 3. Note that the total volume of the simulations is much larger than the survey volume parameters listed in SE03 so that we can study the effect of nonlinearity with little interference from statistical variance. The resulting density field of each simulation box is Fourier transformed, and the squared complex norms of Fourier coefficients are spherically averaged over all simu- lations to give the matter power spectra in the wavenumber- shells of widths ∆k = 0.005h Mpc−1. A mode is in- cluded in a shell if its discrete wavenumber falls in the shell. Because each mode in the discrete transform is in- cluded in one and only one wavenumber-shell, the shells are not correlated for a Gaussian field even though the ∆k is smaller than the size of an independent cell in Fourier space, 2π/V 1/3 box . Using small ∆k ensures that the contrast of narrow features are not artificially reduced. However, because a thin shell contains fewer modes, we do want to apply some smoothing. We use Savitzky-Golay filtering (Press et al. 1992), as this can preserve peak heights bet- ter than boxcar smoothing. The Savitzky-Golay method also gives estimates of the derivatives1. From compar- isons between power spectra before and after smoothing, we believe that this procedure neither introduce mislead- ing baryonic features nor erase meaningful features. For power spectra in redshift space, we spherically average the Fourier transform of the density fields after displacing par- ticles according to their peculiar velocities assuming a dis- tant observer. 3. THE NONLINEAR MATTER POWER SPECTRUM 3.1. Nonlinear effects in the matter power spectrum As mass perturbations on a given scale approach or- der unity in amplitude, linear perturbation theory breaks down and the gravitational growth of perturbations in one mode is increasingly coupled with perturbations in other modes. The higher-order contribution resulting from this mode coupling hinders the detection of features in the ini- tial power spectrum, including the baryonic acoustic os- cillations. That is, the additional power contributed from other modes blurs the initial features at a given Fourier mode as they mix with a convolution of other modes (Jain & Bertschinger 1994, and references therein). Nonlinear growth from mode coupling increases power above the lin- ear growth rate on small scales, resulting in a bigger sta- tistical variance for any underlying initial features. As the amplitude of density perturbations grows with time, these nonlinear effects become stronger and proceed to larger scales. As a basic model, one might distinguish the linear regime from nonlinear regime with a scale R (= π/2kmax) for an appropriate rms overdensity fluctuation, σR, at a given epoch, and assume scales with a smaller rms over- density as linear. In studies of the statistical expectations from large redshift surveys, imposing more conservative values of R improves the linear approximation, but this is at the expense of more of the remnant linear information in the quasilinear regime beyond R. On the other hand, more forgiving criteria for R will result in an overestima- tion of the performance: the precision of the acoustic scale measurement improves as kmax increases, finally saturat- ing beyond about 0.25h Mpc−1 because of Silk damping (see Figure 4 of SE03). Only an N-body study can say whether this nonlinear scale accurately accounts for the erasure of features in the initial power spectrum such as baryonic oscillations. This will give an additional handle in the transition from linear to nonlinear regime beyond that traced by the increased amplitude. We will use the notation kmax in this paper to characterize the nonlinear 1We use 4th order polynomial with a filter width of ∆k = 0.04h Mpc−1 to smooth power spectra and ∆k = 0.05h Mpc−1 to derive the derivatives of power spectra. 3 Fig. 1. -- The matter power spectra at various redshifts. Left: each power spectrum is divided by a zero-baryon power spectrum at the given redshift. Solid lines are for the real-space clustering, and dashed lines with the same color are for the corresponding redshift-space clustering. Right: d ln P/d ln k from the matter power spectra at various redshifts in real space (upper panel) and in redshift space (lower panel). Green dashed line: the input power spectrum from the CMBfast, black line: the linear matter power spectrum at z = 49 generated from the input power spectrum, red: the nonlinear matter power spectra at z = 3, blue: z = 1 and violet: z = 0.3. The first troughs in d ln P/d ln k from the N-body results are lower than that of the input power spectrum due to the interaction of the Savitzky-Golay smoothing with the boundary at k ∼ 0. scale and to assess the linear information from the bary- onic features surviving nonlinearity. Figure 1 shows the effect of nonlinear gravitational growth on baryonic acoustic oscillations in matter power spectra from N-body simulations. The solid lines in the left panel of the figure show the spherically averaged real- space power spectra divided by a zero-baryon power spec- trum2 (Eisenstein & Hu 1998a) at various redshifts. The growth rate calculated from rms overdensity fluctuations at 16h−1 Mpc, σ16h−1 Mpc, is consistent with the linear growth rate, while the values of σ8h−1 Mpc show nonlinear effects at lower redshift. The vertical lines denote the non- linear scales, kmax, adopted in SE03 to satisfy σR ∼ 0.5 (kmax = 0.11h Mpc−1 at z = 0.3, 0.19h Mpc−1 at z = 1 and 0.53h Mpc−1 at z = 3). As expected, nonlinear structure formation increases small-scale power and obscures small-scale baryonic fea- tures. The effect proceeds to larger scales with time. The difference in slope between the linear power spectrum at z = 49 and the power spectra at lower redshift shows that remnant nonlinearity exists even for k < kmax. At lower redshifts, the contrast of baryonic peaks on large scales is decreased because of the nonlinear mode-coupling. The right panel of Figure 1 shows logarithmic derivatives of the matter power spectrum with respect to wavenum- bers, generated from Savitzky-Golay filtering. This plot of derivatives is useful not only because it effectively manifests fine details of power spectra but also because d ln P/d ln k is what enters into the Fisher information ma- trix and creates the standard ruler test with baryonic os- cillations. Of course, the derivative d ln P/d ln k increases the noise in the power spectra despite our use of smooth- 2A power spectrum generated from the fitting formula in Eisen- stein & Hu (1998a) for our fiducial Ωm and h but Ωb = 0 ing. The real-space d ln P/d ln k in Figure 1 (upper panel) demonstrates that the oscillatory features are well distin- guished for k < kmax at z = 0.3 and z = 1 despite the slight nonlinearity. At z = 3, it becomes difficult to distin- guish baryonic features beyond k ∼ 0.4h Mpc−1 because the decreased contrasts of small-scale peaks caused by Silk damping (Silk 1968) make them susceptible to even small amount of nonlinearity or noise. Fortunately, baryonic features in k > 0.3h Mpc−1 have minor contributions to cosmological information (SE03). 3.2. The effect of redshift distortions Since we measure redshifts of galaxies rather than their physical distances, the three-dimensional galaxy power spectrum is subject to redshift distortions, the angle- dependent distortion in power spectra caused by the pe- culiar velocity of galaxies (Kaiser 1987; Hamilton 1998; Scoccimarro 2004). On large scales, the bulk motions of large-scale structure toward overdense regions enhance power, and on small scales, the virial motions within and among halos create an apparent extension along the line of sight, known as the finger-of-God effect (de Lapparent et al. 1986). This suppresses power on small scales. In linear theory, the large-scale power enhancement by the redshift distortions follows a simple form (Kaiser 1987). This is true only for the asymptotic limit of large scale, and in general, the nonlinear effect in velocity fields deviates the redshift-space power from Kaiser formula even on fairly large scales (Scoccimarro 2004). We seek to estimate how much the nonlinear effect of large-scale redshift distortions affects the baryonic features. The dashed lines in the left panel of Figure 1 show the redshift-space power spectra divided by a zero-baryon 4 Fig. 2. -- Left: d ln P/d ln k from Pnonlinear − fL(c0, k, k2) at various redshifts in real space. Gray solid line: the linear matter power spectrum at z = 49, black solid lines: the nonlinear matter power spectra at various redshifts. Right: the redshift-space d ln P/d ln k from Pnonlinear − fL(c0, k, k2) after corrected for the finger-of-God suppression. Note that the gray lines in this panel denote the linear matter power spectrum in real space at z = 49 while black lines are for the nonlinear matter power spectra in redshift space. The vertical dashed lines denote the value of kmax we assumed in SE03. We conclude that the baryonic features survive on large scales despite the nonlinear growth with the redshift distortions imposing an additional but mild degradation. power spectrum. The figure depicts the progress of redshift distortions with nonlinearity and their effect on the bary- onic features in the matter power spectrum. The redshift- space power spectra on large scales have a higher ampli- tude than that of real space by the amount predicted by linear theory (Kaiser 1987) for the asymptotic limit. Even on large scales, we observe a slight suppression in redshift- space power with respect to linear theory, in agreement with Scoccimarro (2004). On small scales, the finger- of-God effect not only suppresses the overall power but also decreases the contrast in baryonic features. As ex- pected, the finger-of-God effect increases with time, and the resulting suppression makes the position of the second peak (which is beyond the nonlinear scale for z = 0.3) appear slightly shifted. For our cases of study, the finger- of-God effect generally produces 10 − 20% of suppression in redshift-space power at kmax in all redshifts when com- pared to the prediction by linear theory. The lower right panel of Figure 1 gives a more clear view of the redshift distortions decreasing the contrasts and in- troducing noise at k < kmax. Nevertheless, the baryonic features are still preserved up to k ∼ kmax for z = 0.3 and z = 1, and k . 0.3h Mpc−1 for z = 3. The result at z = 0.3 seems especially encouraging in that the features are preserved even beyond kmax. It is important to note that these curves are spherically averaged power spectra in redshift space while we aim to use anisotropic information of power spectra in real ob- servations. In the three-dimensional power spectra in red- shift space, wavevectors nearly perpendicular to the line- of-sight direction will preserve baryonic features as well as the real-space power in Figure 1, and wavevectors nearly along the line-of-sight direction will appear more smeared than the redshift-space power in the figure. 3.3. Acoustic features after restoration of the broadband shape In this section, we consider the effect of nonlinear mod- ification of the slope of the power spectrum on the bary- onic features. While the nonlinear mode-coupling effect directly erases the baryonic features, the accompanied in- crease in small-scale power due to nonlinear growth will modify the slope of the power spectrum as a function of wavenumber. The resulting change in slope will shift the apparent locations of the baryonic peaks. A careful look at Figure 1 reveals that the higher harmonics of the baryonic peaks appear at slightly larger wavenumbers compared to the initial power spectrum. Furthermore this addition of broadband power can misleadingly decrease the contrast of the peaks and thereby overestimate the loss in information by the mode coupling effect on the baryonic features. We attempt to correct for the nonlinear growth effect on the broadband shape of the power spectrum by mod- eling the gradual modification to the broadband shape due to nonlinearity as a smooth function of wavenum- ber and subtracting off this smooth function from the nonlinear power spectrum to restore the oscillatory por- tion to its original slope. We fit the nonlinear power spectra in real space to a multiple of the linear power spectrum, g2Plinear, and a 2nd-order polynomial function fL = c0 + c1 k + c2 k2 (with g2 and the ci as constants) and then subtract the smooth function fL from the non- linear power spectra before calculating d ln P/d ln k. This process increases the fractional variation in P by decreas- ing the overall amplitude. In other words, noting that d ln P/d ln k = (1/P ) × (dP/d ln k), we have decreased P but not substantially changed its derivative. Increasing the degree of the function fL up to 3rd-order does not have a sizable effect on d ln P/d ln k on large scales. As required, the resulting function fL is nearly constant on large scales in all cases. The returned values of g2 are not consistent with linear growth rate at lower redshift because the fitting process tends to decrease g2 to account for the erasure of baryonic features. Since we do not want to over- amplify the baryonic features by oversubtracting beyond a true nonlinear power, as a sanity check, we calculate and subtract fL by fixing g2 to the analytic linear growth rate. This reduces the contrast of resulting baryonic features in d ln P/d ln k at z = 0.3 but only by a small amount. The left panel of Figure 2 shows the resulting nonlinear d ln P/d ln k in real space (black lines) in comparison to the linear power spectrum at z = 49 (gray lines). The agree- ment between the power at z = 3 and the linear power spectrum is excellent. We also recover larger contrasts in z = 0.3 and z = 1 cases after the broadband shape is restored. Our result shows that it is possible to trace baryonic features up to k ∼ 0.3h Mpc−1 at z = 1 and k ∼ 0.2h Mpc−1 at z = 0.3. The results suggest that the performance of baryonic features as a standard ruler will be diminished due to the decreased contrast at k < kmax, but the features preserved beyond kmax will tend to com- pensate for the reduction. In short, we find that correcting for the nonlinear effects on the broadband shape with a smooth function helps us rescue baryonic features by some degree, implying that the nonlinear effects are relatively smooth in power. We interpret the amount we cannot recover as having been lost to mode coupling effects. For the redshift-space power spectrum, we first correct for the finger-of-God suppression. We fit the nonlinear power spectrum in redshift space to the counterpart in real-space by a multiplicative smooth function in the form of Ffog = 1/(kmσm + 1)1/m where σ and m are fitting parameters. We find m ∼ 2.3 − 2.95. After dividing by Ffog we calculate and subtract fL to match the restored redshift-space power spectrum to the linear power spec- trum. Although Ffog may not necessarily match the con- ventional form of an exponential finger-of-God suppres- sion, the function behaves properly at the limits of small and large wavenumber and is capable of characterizing the difference between the real-space and redshift-space power from our N-body results. The right panel of Figure 2 shows the resulting d ln P/d ln k in redshift space (black lines) in comparison to the linear power spectrum at z=49 in real space (gray lines). The restored redshift-space power spectra show slightly larger contrasts in baryonic features at lower redshift relative to the uncorrected ones (lower right panel of Figure 1). From the comparison between the real-space and redshift-space d ln P/d ln k (Figure 2), the redshift-space power reasonably traces the details in the real-space power on linear and quasilinear scales al- though the contrast of the baryonic features degrades in redshift space. In summary, the real-space (redshift-space) matter power spectrum traces baryonic features up to k ∼ 0.4 (0.3)h Mpc−1 at z = 3. At z = 1, kmax of 0.19h Mpc−1 seems a rea- sonable choice for the linear approximation in real space considering the features preserved on even smaller scales, but a lenient standard in redshift space. At z = 0.3, kmax ∼ 0.11h Mpc−1 is conservative both in real space and redshift space. Adopting a slightly larger kmax ap- 5 pears justified at this redshift. 4. THE EFFECT OF BIAS 4.1. Anomalous power In galaxy redshift surveys, we do not directly observe the real-space matter power spectrum but instead observe the distribution of biased tracers of matter in redshift space. The assumptions of local bias and Gaussian statistics for the density fields lead to a scale-independent bias on large scales for the correlation function (Coles 1993; Scherrer & Weinberg 1998; Meiksin, White, & Peacock 1999; Coles et al. 1999). Any excess small-scale correlation from biasing will appear as an additional constant term in the biased power spectrum on large scales. This holds even when the matter density field is nonlinear (Scherrer & Weinberg 1998; Coles et al. 1999; Seljak 2000). The bias on large scales is thus scale-independent up to this constant term. On smaller scales, the bias will generically deviate from the simple approximation of scale independence. We use the term 'nonlinear bias effect' in this paper to designate any deviation from a simple multiplicative bias in the power spectrum. As the biased tracers such as galaxies are rare objects compared to the underlying matter, the biased power spec- trum is subject to a shot noise. Conventionally, one writes the shot noise as a white noise (P ∼ constant) equal to the inverse of the number density of particles. However, the effects of the limited number of particles can be more complicated than a simple Poisson noise, such as in bi- asing schemes where halos are not equally weighted or, more enigmatically, the discreteness effects in the process of finding halos. For example, the fact that a halo finder cannot identify two halos arbitrarily close together means that the shot noise will not be white. Therefore it is not straightforward to identify the shot noise with the con- ventional white noise based on an inverse number density. Rather than singling out the shot noise term, we hereafter group this term with the nonlinear bias term without dis- tinguishing one from the other. Both terms are easily ap- proximated as an additive constant on large scales. Here- after, 'anomalous power' is used to refer to a combination of these two terms. We find that the fractional level of the anomalous power compared to the linearly biased power is important to track. We hereafter will describe this frac- tional level as 'larger' or 'smaller' anomalous power. The anomalous power will contribute additional power above the linearly biased power spectrum, therefore in- creasing the statistical variance of the underlying features. The nonlinear bias effect in the anomalous power may also induce a mixture of information from different Fourier modes and so erase features. 4.2. Bias schemes In generating biased tracers of the matter, we do not attempt to reproduce realistic galaxy populations but in- stead to reflect an interesting range of models of galaxy populations using simple deterministic halo-based biasing schemes. The current halo occupation distribution (HOD) models suggest a single galaxy per halo for low-mass halos above a mass threshold and an additional power-law mean occupation for more massive halos while the details vary in different studies of galaxy populations (Berlind et al. 6 Fig. 3. -- Biased power spectra at z = 1 divided by a zero-baryon power spectrum. The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). The label 'm' denotes the minimum group multiplicity for selecting halos. For the NUM cases, b ∼ 1.7, 2, and 2.5, and for the MASS cases, b ∼ 2.4, 2.7, and 3.1 as m increases. The solid lines are for the real-space power, and the dashed lines are for the redshift-space power. 2003; Kravtsov et al. 2004; Zehavi et al. 2004). We reduce the complexity by decomposing the HOD models into the two extreme bounding cases: one in which halos above a mass threshold host a single galaxy and one in which ha- los above a mass threshold have mean occupation as linear to the halo mass (i.e., the power-law index of unity). To create different amplitudes of bias, we apply various min- imum group multiplicities (i.e., halo mass-thresholds) for both cases. Various superpositions of these trial cases then can comprise more complex HOD models. Note that the resulting populations themselves already represent super- positions of different-mass halos. Therefore our biasing schemes will be sufficient for examining the robustness of the baryonic features in various galaxy populations be- cause, if baryonic oscillations are found to survive in both of the extreme cases, it seems unlike that a mixture would fair worse. This will be further justified once our results can show that different bias models indeed extrapolate to linear biasing on large scales. We use the friends-of-friends method (Davis et al. 1985) and identify halos by adopting a linking length of 0.6 h−1 Mpc and minimum group multiplicities of 4, 10, or 30 particles. We assign zero galaxy density for the re- gions below this threshold. For the regions identified as halos with a given threshold, we assign galaxies with two schemes: 1. Number Weighted (NUM), where we assign one galaxy per halo, using the center-of-mass position and ve- locity. In this case, the number of galaxies does not follow the mass of halos, and the information on the virialized motions within each halo (the finger-of-God effect) is lost. 2 Mass Weighted (MASS), where we retain the velocity and density structure of the halo by assigning one galaxy per simulation particle. This way, the number of galaxies follows the mass of the halo, and the finger-of-God effect is preserved. The minimum group multiplicity, m, of 4 and 10 are small compared to 20 − 30 particles usually desired for robust halo identification (Somerville et al. 2000; Jenk- ins et al. 2001). However, the low multiplicity halos are still tracing overdense regions, albeit in a more stochas- tic manner. This randomness would not be favorable for recovering acoustic oscillations, implying that our results are conservative. At lower redshifts, both methods are able to provide us a sufficient range of bias values of our interest. However, at z = 3, the mass resolution of our simulation is too low to find halos with masses low enough to yield mild values of biases (i.e. b ∼ 3). For this reason, we use an addi- tional bias scheme at this redshift, where the density of the tracers approximately follows the matter density squared (hereafter, Rho2). In this biasing scheme, each particle is weighted by the average density with a 2h−1 Mpc spline- smoothing kernel centered on the particle. We implement this with the SMOOTH code3. From the distribution of the weighted particles, we calculate the density in each mesh in real space and redshift space. The finger-of-God effect is preserved in redshift space in this scheme. We first discuss the biased power spectra at z = 1 and z = 0.3 because the two share common bias schemes. We then present the z = 3 results. 4.3. Bias effects at z = 1 Figure 3 shows power spectra of biased tracers divided by a zero-baryon power spectrum in real space and redshift space. The left and right panels show number-weighted cases (NUM) and mass-weighted cases (MASS), respec- tively, with the minimum group multiplicity m, of 4, 10, and 30. NUM cases with m = 4, 10, and 30 produce biased tracers with b ∼ 1.7, 2, and 2.5 where b is the ratio of 3http://www-hpcc.astro.washington.edu/tools/ 7 Fig. 4. -- d ln P/d ln k from the biased power spectra at z = 1 in real space (upper panels) and in redshift space (lower panels). The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). Green dashed line: the input power spectrum, black: the nonlinear matter power spectrum at z = 1, red: biased with m = 4, blue: biased with m = 10 and violet: biased with m = 30. the biased power spectra to the matter power spectra on large scales. The mass-weighted tracers (MASS) in the right panel of Figure 3 show larger biases than number- weighted cases for the same m: b ∼ 2.4, 2.7, and 3.1. This is because the mass-weighted cases give more weight to the high-mass halos with larger biases while the number- weighted cases are dominated by halos close to the mass threshold. In both cases, the relative anomalous power increases as the amplitude of bias increases. For example, the m = 4 cases show little anomalous power, and the m = 10 and m = 30 cases show larger but still mild anoma- lous power for k < kmax (0.19h Mpc−1). When different bias models with the same multiplicity are compared, the MASS cases have slightly larger anomalous power up to k ∼ 0.3h Mpc−1. When different bias models with a simi- lar value of bias are compared (e.g., m = 4 of MASS and m = 30 of NUM), the MASS case exhibits smaller anoma- lous power up to k ∼ 0.3h Mpc−1 than the NUM case. MASS cases with all galaxies placed at the center of a halo showed the same trend, which means that the trend is not due to the effect of the halo density-profile. Thus, the bias scheme for the NUM cases introduces larger anomalous power, be it from the shot noise or from the nonlinear bias effect. One interesting question will be whether the differ- ence in anomalous power directly relates to the erasure of baryonic features. The upper panels in Figure 4 show d ln P/d ln k of biased power in real space (solid lines) in comparison to the input power spectrum (dashed lines). In the figure, all biased power spectra in real space preserve oscillatory features at least up to k ∼ 0.2h Mpc−1 while the contrast appears decreasing due to the nonlinear bias effect as bias increases. As before (§ 3.3), adding a smooth component due to a shot noise or nonlinear bias to the power spectrum will decrease the contrast in baryonic features. Subtracting off this smooth component will help to recover some of the baryonic features. We will revisit this in § 4.4. In redshift space (dashed lines in Figure 3), the power spectra of the MASS cases clearly show the finger-of-God suppression as k increases. The finger-of-God suppression is mostly removed in the spectra of the NUM cases, as is to be expected from the methods of biasing, and they show a mild remnant suppression in power with respect to the Kaiser formula4 as k increases. The lower panels of Figure 4 show that baryonic oscillations are smeared more in redshift space than in real space not only in MASS cases but also in NUM cases despite the suppressed finger-of- God effect in the latter case. Again, until we remove the bias effect on the broadband shape, it is hard to determine the degree of erasure. 4.4. Bias effects at z = 1 with the broadband shape restored We next subtract the anomalous power from the biased power spectrum to restore the broadband shape and elim- inate the superficial decrease in contrast of the baryonic features. The restoring process we adopt is intended to assess the optimal amount of information on baryonic fea- tures available from the biased tracers. In real galaxy red- shift surveys, this unbiasing would be determined simul- taneously in the parameter estimation, which is an addi- tional complication. We fit the biased power spectra to a multiple of the non- linear matter power spectrum b2Pmatter at the given red- shift plus a polynomial function fNL = c0+c1 k+c2 k2. We then subtract the smooth function fNL that represents the anomalous power from the biased power spectra when cal- culating d ln P/d ln k. As the anomalous power from bias 4Kaiser formula in our context means (1 + 2β(k)/3 + β2(k)/5) m (z)/b(k), and b(k) is calculated from the ratio where β(k) = Ω0.6 between the biased power spectra and the matter power spectra. 8 Fig. 5. -- d ln P/d ln k from Pbiased − fNL(c0, k, k2) at z = 1 in real space. The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). The label m denotes the minimum group multiplicity for selecting halos. Gray dashed line: the input power spectrum, gray solid line: the nonlinear matter power spectrum at z = 1, black solid lines: the biased power spectra at z = 1. The vertical dashed line denotes kmax = 0.19h Mpc−1. One sees that the biased power spectra follow the features in the underlying matter power spectrum fairly well on linear and quasilinear scales once the broadband shape is restored. is smooth, this will help recover baryonic features. The ranges of wavenumber in real or redshift space is chosen suitably at each redshift so as to span well beyond the lin- ear region to constrain fNL but short enough so as not to weight the fit too much towards large wavenumbers. Vari- ations in fNL due to choosing different ranges of wavenum- bers or different degrees of the function (up to 3rd-order) do not show a meaningful impact on d ln P/d ln k on large scales. The resulting function fNL behaves as constant on large scales as required, although on small scales, fNL for the MASS cases show a slow roll-over due to the extended halo profiles compared to the NUM cases. Figure 5 shows the resulting derivatives at z = 1 af- ter anomalous power fNL is subtracted (black lines) in comparison to the nonlinear (solid gray lines) and input matter power spectra (dashed gray lines). For k < kmax, the agreement between the biased power spectra and the nonlinear matter power spectrum is excellent regardless of the different biasing schemes. Beyond kmax, we observe small variations depending on bias schemes. The varia- tions appear related to the amount of anomalous power as this contributes additional power above the underlying baryonic features, increasing the statistical noise. Within the same bias models, higher mass thresholds and hence larger biases yield noisier derivatives, as would be expected from the increase in anomalous power. If we compare two bias models with similar bias values, m = 4 of MASS and m = 30 of NUM, then the baryonic features in m = 30 of NUM appear noisier as it has a larger anomalous power. However, when those with the same group multiplicity are compared between different bias models, the baryonic fea- tures in the MASS cases look no worse than those in the NUM cases even with the larger bias and the slightly larger anomalous power for k . 0.3h Mpc−1. Despite slight variations depending on bias schemes, the baryonic features in general have not been obviously dam- aged by the biasing process; the biased power spectrum closely follows the details in the underlying nonlinear mat- ter power spectrum over a broad range beyond kmax. This is different from the effect of nonlinear gravity shown in § 3.3 where we saw the effect of mode coupling. This sup- ports the hypothesis that the anomalous power, whether due to shot noise or nonlinear bias, is a smooth function of wavenumber. The contrasts of the baryonic features appear slightly larger in some of the biased power spectra relative to the matter power spectrum, but this is likely due simply to the increased noise. For the redshift-space power spectrum of the MASS cases, we apply a similar fitting process to the one in § 3.3 to restore the broadband shape both from the nonlinear bias and nonlinear redshift distortions. First we correct for the finger-of-God effect with a multiplicative function Ffog to match the biased power in redshift space to the biased power in real space. We then calculate and subtract fNL to remove the anomalous power. For the redshift-space power spectrum of the NUM cases, we do not correct for the finger-of-God suppression despite the slight deviation from the Kaiser formula. We calculate and subtract an additive fNL from the redshift-space power in this case. Figure 6 shows the resulting derivatives at z = 1 in redshift space (black lines) in comparison to the non- linear matter power spectrum in real space (gray lines). In redshift space, the contrast of the last feature before k = 0.2h Mpc−1 is smaller than in real space but still in good agreement. Beyond kmax, we see the traces of bary- onic features although they look noisier than in real space. Again, the nonlinear effect of redshift distortions on bary- onic features in the NUM cases are no better than that in the MASS cases even though the virialized motions within the halos are suppressed in NUM cases, and this probably is related to the nonlinear effect on the velocity fields on large scales (Scoccimarro 2004). 9 Fig. 6. -- d ln P/d ln k from Pbiased − fNL(c0, k, k2) at z = 1 in redshift space. Note that the MASS cases are corrected for the finger-of-God suppression beforehand. The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). Gray: the nonlinear matter power spectrum at z = 1 in real space, black: the nonlinear matter power spectrum in redshift space (NO BIAS) and biased power spectra in redshift space. Note that the matter power spectrum in redshift space (NO BIAS) is also corrected with Ffog to fit to the matter power in real space. The fitting range (k < kfit) is indicated by the extent of the black lines. The number-weighted cases are fitted to kfit = 0.5h Mpc−1 but kfit = 0.35h Mpc−1 elsewhere. The vertical dashed line denotes the value of kmax. The redshift-space biased power spectrum nearly reproduces the baryonic features of the real-space biased power spectrum with mild degradation. 4.5. Bias effects at z = 0.3 We next investigate the effects of bias at lower redshift. Figure 7 shows power spectra of biased tracers divided by a zero-baryon power spectrum at z = 0.3. The NUM cases (left panel) generate tracers with b ∼ 1.2, 1.3, and 1.6, and the MASS cases (right panel) generate b ∼ 1.8, 2, and 2.3. At this redshift, we are particularly interested in tracers with b ∼ 2, which corresponds to the lumi- nous red galaxy sample (LRG) of Sloan Digital Sky Survey (SDSS). This corresponds to m = 10 in the MASS cases although anomalous power in the m = 10 case is half the inverse of the number density of galaxies in the LRG sam- ple, suggesting that the m = 10 bias model is not exactly right. The bias and relative anomalous power is small up to k ∼ 0.15h Mpc−1 in all cases relative to z = 1. Recov- ering baryonic features beyond kmax (0.11h Mpc−1) will be possible both in real space and in redshift space from Figure 8. Figure 9 shows the derivatives in real space after a corresponding smooth function fNL is subtracted. From the figure, the baryonic features in biased power spec- tra trace those in the matter power spectrum well up to k ∼ 0.2h Mpc−1. Again, despite the larger biases of the MASS cases, they preserve baryonic features no worse than the NUM cases. Figure 10 shows the derivatives in red- shift space compared to the matter power spectrum in real space. All biased redshift-space power spectra trace the features in the real space matter power spectrum fairly well up to k ∼ 0.2h Mpc−1 but with more degradation in the NUM cases. To summarize, subtracting the smooth anomalous power helps to recover the contrast of the baryonic features at z = 0.3 as well. With small biases (b . 2), the recov- ered contrast is comparable to the contrast in underlying matter power spectra even in quasilinear scales meaning that the nonlinear scales deduced in § 3.3 is valid despite biasing. Also the recovered contrasts do not seem very sen- sitive to moderate variations of biases, which is consistent with the results at z = 1. 4.6. Bias effects at z = 3 We next show the results at our highest redshift bin, z = 3. The number of simulation boxes used for this red- shift is 30, which is smaller than the other redshift bins. We generated three biased tracers: m = 4 for MASS, m = 4 for NUM, and Rho2 (Figure 11). The former two cases generate power spectra with b ∼ 5.5 and 4.9, which are too high for Lyman break galaxies (Steidel et al. 1996). Correspondingly, the number density of these halos is very small, leading to significant noise in the power spectra. The Rho2 model, on the other hand, generates a bias of b ∼ 2.5, similar to that of Lyman break galaxies (LBG). The anomalous power of Rho2 is 70% of the shot noise ef- fect from the number density of 10−3h3 Mpc−3 that we as- sumed for the sample in SE03 although the relative effect is nearly equivalent. The relative anomalous power amplifies the power by a factor of two at k ∼ kmax(= 0.53h Mpc−1) for the Rho2 case. While the matter power spectrum preserves baryonic peaks up to k ∼ 0.4h Mpc−1 (Figure 1), Figure 12 shows that even the Rho2 case cannot probe baryonic features beyond k ∼ 0.35h Mpc−1 either in real space (left panel) or redshift space (right panel). The MASS and NUM cases trace the acoustic oscillations only up to k ∼ 0.25h Mpc−1. That is, the tracers with very large bias do not mimic the underlying matter power spectra very well, unlike tracers with moderate biases at lower redshift bins. This is likely 10 Fig. 7. -- Biased power spectra at z = 0.3 divided by a zero-baryon power spectrum. The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). The label m denotes the minimum group multiplicity. For the NUM cases, b ∼ 1.2, 1.3, and 1.6, and for the MASS cases, b ∼ 1.8, 2, and 2.3 as m increases. The solid lines are for the real-space power, and the dashed lines are for the redshift-space power. Fig. 8. -- d ln P/d ln k from the biased power spectra at z = 0.3 in real space (upper panels) and redshift space (lower panels). The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). Green dashed line: the input power spectrum, black: the nonlinear matter power spectrum at z = 0.3, red: biased with m = 4, blue: biased with m = 10 and violet: biased with m = 30. 11 Fig. 9. -- d ln P/d ln k from Pbiased − fNL(c0, k, k2) at z = 0.3 in real space. The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). Gray dashed line: the input power spectrum, gray solid line: the nonlinear matter power spectrum at z = 0.3, black solid lines: the biased power spectra. Fitting to kfit = 0.5h Mpc−1. The vertical dashed line denotes kmax = 0.11h Mpc−1. Fig. 10. -- d ln P/d ln k from Pbiased − fNL(c0, k, k2) at z = 0.3 in redshift space. Note that the MASS cases are corrected for the finger- of-God suppression beforehand. The left panel is for the number-weighted cases (NUM), and the right panel is for the mass-weighted cases (MASS). Gray: the nonlinear matter power spectrum in real space, black: the nonlinear matter power spectrum in redshift space (NO BIAS) and biased power spectra in redshift space. The fitting range (k < kfit) is indicated by the extent of the black lines. For the number-weighted cases (NUM), kfit = 0.5h Mpc−1 but kfit = 0.35h Mpc−1 elsewhere. The vertical dashed line denotes the value of kmax. 12 Fig. 11. -- Biased power spectra at z = 3. Left: biased power divided by a zero-baryon power spectrum. Solid lines are for the real space clustering, and dashed lines with the same color are for the corresponding redshift-space clustering. Right: d ln P/d ln k from the biased power spectra. Green dashed line: the input power spectrum, black: the matter power spectrum at z = 3, red: Rho2, blue: NUM with m = 4 and violet: MASS with m = 4. because of the statistical noise from a small number of high mass halos, but we cannot exclude the possibility of an emergence of mode coupling as the nonlinear bias effect becomes very large. We estimate that galaxies with b ∼ 3 at this redshift will recover baryonic features up to wavenumber of about 0.3h Mpc−1. The details of the result may vary depending on the biasing schemes and the shot noise. 4.7. Summary of the effect of bias To summarize the general effect of bias on baryonic fea- tures, subtracting a smooth function to match the slope of the matter power spectrum largely restores the bary- onic features in the underlying matter power spectra if the amplitude of bias is moderate. This implies that a mod- erate anomalous power of bias whether from shot noise or a nonlinear bias effect does not erase the initial fea- tures. In other words, it is reasonable to assume that the anomalous power is smooth in wavenumber and does not generate features that mimic baryonic oscillations. We find that the detailed effects of bias not only scale with the amplitude of bias and the anomalous power but also depend on the biasing schemes used. For very large amplitudes of bias, we clearly lose information whether it is due to increased nonlinear bias effects or merely shot noise. Based on the baryonic features preserved in the matter power spectra (§ 3) and the effect of bias, we summa- rize the nonlinear scales in the biased power spectra. At z = 0.3, the results are encouraging in that moderately biased power spectra (b . 2) preserve baryonic features at k < kmax (0.11h Mpc−1) and record a fair amount of lin- ear information even beyond kmax in both real and redshift space. The biased power spectra at z = 1 (b . 2) show de- creased contrast for k < kmax (0.19h Mpc−1) and contain attenuated traces of baryonic features beyond kmax. At z = 3, the biased power spectrum with b ∼ 3 will preserve the features up to k ∼ 0.3h Mpc−1, but no further. The biased power in redshift space traces the real-space biased power reasonably well with partial degradations in the baryonic features depending on biasing schemes. Sup- pressing the finger-of-God effect does not help to preserve real-space features any better, implying that the motions between halos do not strictly respect linear theory. To this point, we have shown the successful restoration of baryonic features in linear and quasilinear scales from various biasing schemes based on halo-mass thresholds. Given that all mass thresholds preserve the features, it is unlikely that more complex descriptions of galaxy pop- ulations, e.g., superpositions of our biasing schemes, would change the results. Similarly, a more stochastic local bias would not likely remove the oscillations, although stochas- tic models can increase anomalous power (Dekel & Lahav 1999; Scherrer & Weinberg 1998) thereby reducing con- trast and increasing noise. 5. IMPACTS ON COSMOLOGICAL DISTANCE ESTIMATION : χ2 ANALYSIS OF THE N-BODY DATA We next consider the impact of nonlinearity and bias on the statistical constraints on cosmological distances from the baryonic features. We are interested in using the power spectrum measurement to constrain a distance scale, and therefore we want to estimate how well we can constrain dilations in wavenumbers. We define this dilation parame- ter as α(= kref /ktrue). The error on the dilation parameter α represents the errors on the angular diameter distance DA(z) and Hubble parameter H(z). We perform a χ2 anal- ysis to fit the spherically averaged power spectra in real space from N-body simulations, Pobs, to a linear combi- nation of the input linear power spectrum, Plinear, and an additional polynomial function (eq. [1]). The fit parame- ters are α, a multiplicative bias b0, a scale-dependent bias b1, and additive terms for nonlinear growth or an addi- tional constant (a0, a1 and a2). The mean value of α is 13 Fig. 12. -- d ln P/d ln k from Pbiased − fNL(c0, k, k2) at z = 3 in real space and redshift space. The Rho2 case and the MASS case in redshift space are corrected for the finger-of-God suppression beforehand. Left: the real-space power spectra. Gray dashed line: the input power spectrum in real space, gray solid line: the matter power spectrum in real space, black solid lines: the biased power spectra in real space. Right: the redshift-space power spectra. Gray solid line: the matter power spectrum in real space, black solid lines : the biased power spectra in redshift space. Note that the y-axes in the left and the right panels are not scaled the same unlike the previous figures. The vertical line is at kmax of 0.53h Mpc−1. From the figure, the tracers with a very large bias do not mimic the underlying matter power spectra very well. One sees that the mildly biased Rho2 case (b ∼ 2.5) probes baryonic features up to k ∼ 0.35h Mpc−1. expected to be unity since we set the reference cosmology to be equal to the true cosmology for simplicity. ref ) We try two cases, with and without including b1. Pobs(kref ) = (b0+b1kref )×Plinear(kref /α)+(a0+a1kref +a2k2 (1) In both cases, we fit power at wavenumbers beyond kmax to set the nonlinear trends. We note that even though the model spectrum contains higher harmonics of the baryon oscillations that may be absent from the data, this need not bias the distance measurement. The smooth portion of the nonlinear power spectrum that replaces higher har- monics does not have narrowband features to match on the linear model and hence impose a preferred physical scale. We will return to this point later in this section. The mean value and error of α are computed using jack-knife subsampling of the simulations (Lupton 1993). Since we do not know the true covariance matrix, we as- sume a Gaussian error in each k-bin constructed from the power spectrum averaged over all sets. Although assum- ing a Gaussian error implies that we underestimate the statistical noise relative to the true non-Gaussian error, the variations among jack-knife estimates of α in the 51 subsamples (30 at z = 3) should reflect the non-Gaussian, mode-coupled error as these subsamples are drawn from actual nonlinear N-body data. In other words, our fit- ting slightly misweights the data relative to optimum but should not produce overly optimistic σα compared to the true error. The behavior of resulting errors are consistent with the effect of nonlinearity and bias that we studied in the pre- vious sections. For the underlying matter power spectra at z = 3, we derive σα ∼ 0.35(0.36)% with wavenumbers less than kfit = 0.3(0.5)h Mpc−1. For the Rho2 biased power spectrum, σα ∼ 0.35% when kfit ∼ 0.3h Mpc−1. Increasing the range of k beyond this value increases the error, which is consistent with the noisy feature in the left panel of Figure 12. If we scale the error to the survey volume assumed in SE03, this error value corresponds to 1%. The anomalous power in the Rho2 case is close to the shot noise we assumed there. The analytic results in SE03 implies σα ∼ 0.93% (from 1/σ2 H(z)), in good agreement. DA(z) + 1/σ2 α = 1/σ2 At z = 1, we calculate σα ∼ 0.4% for the underlying nonlinear matter power spectrum if we include a region up to kfit = 0.3h Mpc−1. Among the biased power spec- trum, the NUM case with m = 4 has a bias value close to that of the redshift bin at z = 1 in SE03. For this case, σα ∼ 0.4 − 0.5%, which corresponds to σα ∼ 1.6 − 1.8% when scaled to the baseline survey volume in SE03. This is to be compared with σα of 1.4% for the corresponding number density in SE03. Thus the simulation indicates a slightly worse precision relative to our previous predic- tions in SE03. The equivalent kmax for a linear approxi- mation will be 0.17 − 0.18h Mpc−1 at z = 1, rather than 0.19h Mpc−1. At z = 0.3, we find more optimistic results, as we would expect from Figure 9. For the underlying nonlinear matter power spectrum, we get σα ∼ 0.6% for kfit = 0.3h Mpc−1. For the MASS cases with m = 10 and m = 30, which are similar to LRG samples, we find σα ∼ 0.8 − 0.9%. When scaled for a survey volume of 1h−3 Gpc3, σα is 2.1− 2.3% (the equivalent kmax ∼ 0.15 − 0.155h Mpc−1), which is better than the 3.9% calculated from values in SE03. This is to be compared with the current observations: the 4% measurements from Eisenstein et al. (2005) would give σα ∼ 3% when scaled to 1h−3 Gpc3. The cause of the difference between 2.1 − 2.3% and 3% is due to the neglect 14 of redshift distortions in this modeling. In general, the errors calculated with and without b1 are consistent. Without b1, the mean values of α are close to unity, indicating negligible bias. But α is slightly biased above 1 in cases with b1, particularly at lower redshift, albeit by < 1% in most of cases. Since we use the lin- ear power spectrum to match nonlinear power, the fitting process favors a negative b1 to match the erased baryonic features and, in order to compensate the resulting phase shifts of the oscillations, biases α preferentially above 1. On the other hand, without b1, the fitting process does not have means to use bias on α to account for the erased features, giving little bias on α. Including an appropriate recipe to account for the erasure of the baryonic features should remove this bias. That is, this bias would likely be easy to calibrate and remove using N-body simulations of a reasonable cosmological model. The results of χ2 analysis can be translated to the survey volumes required to achieve σα ∼ 1%. Table 1 presents estimates of the required survey volumes, assuming biased power spectra in real space. We do not extend the χ2 analysis to the redshift-space power spectra because of our lack of a reliable model of the redshift distortions to fit. There are deviations from the Kaiser formula on large scales as well as the finger- of-God effect on intermediate and small scales, both of which could involve an arbitrary angular dependence in two dimensions (reduced from three dimensions by the azimuthal symmetry). Nevertheless, the comparisons be- tween the real-space and redshift-space power spectra in our study lead to qualitative estimations of the effects of redshift distortions. Due to the decreased contrast of the baryonic features we observed in the spherically averaged power spectra in redshift space, we expect that the errors on H(z) will be degraded relative to the analytic predic- tion. At z = 3, we expect the effect will be insignificant. At z = 1, the degradation will produce a larger error on H(z) than in SE03. At z = 0.3, a degradation due to the redshift distortions will increase σH but likely no worse than the estimates in SE03. It is important to note that the information in the spherically averaged redshift-space power spectra will not be quantitatively equivalent to that in the two-dimensional redshift-space power spectra. Here we are averaging out the nonlinear redshift distortions, which are actually angle-dependent. For example, a sim- ple exercise of χ2 analysis using only the modes nearly along the line of sight suggests that the degradation of H(z) at z = 1 could be as large as a factor of two between real space and redshift space. We plan to investigate this in a future study. 6. CONCLUSION We have used a large set of N-body simulations to show that the baryonic oscillations from the large galaxy red- shift surveys survive on large scales well despite the mild nonlinearity of gravity, redshift distortions, and bias. We compared the nonlinear effect on the baryonic features ob- served in the N-body results with the choices of nonlinear scale kmax in SE03. As expected, the nonlinear gravitational evolution erased the baryonic features progressively from smaller scales to larger scales as the redshift decreased. In real space, the nonlinear scales we have assumed in SE03 seem fairly con- servative at z = 0.3 and z = 1. We find that we need a slightly smaller kmax for z = 3, but this modification has a minor effect on standard ruler test due to the small nonlin- ear scale. The redshift distortions imposed an additional obscuration for k < kmax. Nevertheless, the redshift-space power spectra reasonably traced the features in the real- space power. We have shown that moderate nonlinear bias (b < 3) does not erase the initial features. The biased power spec- tra follow the features in the underlying matter power spectrum fairly well once the broadband shape is restored. The effect of bias is not only proportional to the ampli- tude of the bias and anomalous power but also depends on the biasing models as well. However, these dependences seem to have minor effect on the underlying baryonic fea- tures. In redshift space, suppressing the finger-of-God ef- fect does not improve recovery of the features, and this indicates that nonlinear effects in the velocity fields on large scales obscure baryonic features as well. Neverthe- less, the redshift-space biased power spectrum reproduces the baryonic features of the real-space counterpart reason- ably well despite the degradation mainly due to the modes along the line of sight. From χ2 analysis of N-body results in real space, we predict errors on cosmological distances similar to those in SE03. Thus the nonlinear scale kmax we have adopted in SE03, with minor modification detailed below, ade- quately describes the effect of nonlinearity on the standard ruler test. Furthermore, this implies that the cosmological distortions will be indeed distinguishable from nonlinear growth and scale-dependent bias, and so the derived un- certainty on cosmological distances depends on the degree of erasure of the baryonic features. Considering nonlinear gravity, redshift distortions and the clustering bias effect from N-body results all together, we estimate nonlinear scales appropriate for calculation of the information in the baryonic features. We consider both the loss of informa- tion for k < kmax and the additional linear information on smaller scales which compensates the loss. Referring to the results from the χ2 analysis, the biased power spectra at z = 3 with b ∼ 3 traces baryonic features well up to k ∼ 0.3h Mpc−1 both in real and redshift space. At z = 1 with b . 2, kmax ∼ 0.17 − 0.18h Mpc−1 are appropriate in real space, but we need a slightly smaller kmax for redshift space. At z = 0.3 with b . 2, kmax ∼ 0.15h Mpc−1 for real space but smaller kmax(> 0.11h Mpc−1) if we consider redshift distortions. We translate these results to the dis- tance measurements: at z = 3 and z = 0.3, we expect that DA(z) will be constrained as well as the estimates in SE03 while H(z) will be slightly less well constrained because of the nonlinear redshift distortion effect on the baryonic features. At z = 1, the deviation will be the largest. While DA(z) will be near the estimates in SE03, H(z) can be as large as twice of DA(z). To summarize, the standard ruler test using baryonic features are robust against nonlinear effects in the lin- ear and quasilinear regime. Therefore, using the standard ruler, the on-going and future large galaxy redshift sur- veys will measure the dilations in observed scales due to cosmological distortions at various redshifts to excellent accuracy, providing a superb probe of the acceleration his- tory of the universe. We thank Martin White for helpful discussions. We were supported by grant AST-0098577 and AST-0407200 from the National Science Foundation. DJE is further sup- ported by an Alfred P. Sloan Research Fellowship. Amendola, L., Quercellini, C., & Giallongo, E. 2005, MNRAS, 357, 429 REFERENCES 15 Angulo, R., Baugh, C. M., Frenk, C. S., Bower, R. G., Jenkins, A., & Morris, S. L. 2005, astro-ph/0504456 Bennett, C. L., et al. 2003, ApJS, 148, 97 Benoit, A. et al. 2003, A&A, 399, L19 Berlind, A. A., et al. 2003, ApJ, 593, 1 de Bernardis, P., et al., 2000, Nature, 404, 955 Blake, C., & Glazebrook, K. 2003, ApJ, 594, 665 Bond, J.R. & Efstathiou, G. 1984, ApJ, 285, L45 Cole, S., et al. 2005, astro-ph/0501174 Coles, P. 1993, MNRAS, 262, 1065 Coles, P., Melott, A.L., & Munshi, D., 1999, ApJ, 521, L5 Cooray, A. 2004, MNRAS, 348, 250 Couchman, H. M. P., Thomas, P. A., & Pearce, F. R. 1995, ApJ, 452, 797 Davis, M., Efstathiou, G., Frenk, C. S., & White, S. D. M. 1985, ApJ, 292, 371 Dekel, A., & Lahav, O. 1999, ApJ, 520, 24 de Lapparent, V., Geller, M. J., & Huchra, J. P. 1986, ApJ, 302, L1 Dolney, D., Jain, B., & Takada, M. 2004, astro-ph/0409445 Eisenstein, D.J., & Hu, W. 1998a, ApJ, 496, 605 Eisenstein, D. J., Hu, W., & Tegmark, M. 1998, ApJ, 504, L57 Eisenstein, D.J., 2003, in Wide-field Multi-Object Spectroscopy, ASP Conference Series, ed. A. Dey Eisenstein, D. J., et al. 2005, astro-ph/0501171 Halverson, N. W. et al. 2002, ApJ, 568, 38 Hamilton, A.J.S., 1998, "The Evolving Universe", ed. D. Hamilton (Kluwer: Dordrecht), p. 185; astro-ph/9708102 Hanany, S., et al., 2000, ApJ, 545, L5 Hatton, S., & Cole, S. 1998, MNRAS, 296, 10 Holtzman, J. A. 1989, ApJS, 71, 1 Hu, W., & Haiman, Z. 2003, Phys. Rev. D, 68, 063004 Hu, W. & Sugiyama, N. 1996, ApJ, 471, 542 Jain, B. & Bertschinger, E. 1994, ApJ, 431, 495 Jenkins, A., Frenk, C. S., White, S. D. M., Colberg, J. M., Cole, S., Evrard, A. E., Couchman, H. M. P., & Yoshida, N. 2001, MNRAS, 321, 372 Kaiser, N. 1987, MNRAS, 227, 1 Kravtsov, A. V., Berlind, A. A., Wechsler, R. H., Klypin, A. A., Gottlober, S., Allgood, B., & Primack, J. R. 2004, ApJ, 609, 35 Lee, A. T., et al. 2001, ApJ, 561, L1 Linder, E. V. 2003, Phys. Rev. D, 68, 083504 Lupton, R. H. 1993, "Statistics in Theory and Practice", (Princeton: Princeton Univ. Press) Matsubara, T. 2004, ApJ, 615, 573 Meiksin, A., White, M., & Peacock, J. A. 1999, MNRAS, 304, 851 Meiksin, A., & White, M. 1999, MNRAS, 308, 1179 Miller, A.D., Caldwell, R., Devlin, M.J., Dorwart, W.B., Herbig, T., Nolta, M.R., Page, L.A., Puchalla, J., Torbet, E., & Tran, H.T., 1999, ApJ, 524, L1 Netterfield, C. B., et al. 2002, ApJ, 571, 604 Pearson, T. J., et al. 2003, ApJ, 591, 556 Peebles, P. J. E. & Yu, J. T. 1970, ApJ, 162, 815 Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1992, "Numerical Recipes in FORTRAN", (Cambridge: Cambridge University Press) Scherrer, R.J., & Weinberg, D.H. 1998, ApJ, 504, 607 Scoccimarro, R., Zaldarriaga, M., & Hui, L. 1999, ApJ, 527, 1 Scoccimarro, R. 2004, Phys. Rev. D, 70, 083007 Seljak, U., & Zaldarriaga, M. 1996, ApJ, 469, 437 Seljak, U., 2000, MNRAS, 318, 203 Seljak, U. 2001, MNRAS, 325, 1359 Seo, H. & Eisenstein, D. J. 2003, ApJ, 598, 720 Silk, J. 1968, ApJ, 151, 459 Somerville, R. S., Lemson, G., Kolatt, T. S., & Dekel, A. 2000, MNRAS, 316, 479 Spergel, D. N., et al. 2003, ApJS, 148, 175 Springel, V., et al. 2005, astro-ph/0504097 Steidel, C. C., Giavalisco, M., Pettini, M., Dickinson, M., & Adelberger, K. L. 1996, ApJ, 462, L17 White, M. 2001, MNRAS, 321, 1 Zaldarriaga, M., Seljak, U., & Bertschinger, E. 1998, ApJ, 494, 491 Zaldarriaga, M., & Seljak, U. 2000, ApJS, 129, 431 Zehavi, I., et al. 2004, astro-ph/0408569 16 Required Survey Volume to Achieve σα = 1% Table 1 Redshift 0.3 1 3 bias 2.0 2.3 1.7 2.4 2.5 Vsurvey(h−3 Gpc3) 4.4 5.5 1.4 1.7 0.5 neff (h3 Mpc−3) 2.1 × 10−4 1.2 × 10−4 2.1 × 10−3 3.4 × 10−4 1.4 × 10−3 Note. -- The approximate survey volumes required to achieve σα = 1% for the biased power spectra. Note that the values are based on our χ2 analysis of the real space power spectra, and this means that redshift distortions are not accounted for. The effective number density neff is the inverse of the anomalous power at k ∼ 0.
astro-ph/9901263
1
9901
1999-01-20T08:14:49
A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies
[ "astro-ph" ]
We present spectroscopic observations of galaxies in the fields of 10 distant clusters for which we have previously presented deep imaging with WFPC2 on board the HST. The clusters span the redshift range z=0.37-0.56 and are the subject of a detailed ground- and space-based study to investigate the evolution of galaxies as a function of environment and epoch. The data presented here include positions, photometry, redshifts, spectral line strengths and classifications for 657 galaxies in the fields of the 10 clusters. The catalog comprises 424 cluster members across the 10 clusters and 233 field galaxies, with detailed morphological information from our WFPC2 images for 204 of the cluster galaxies and 71 in the field. We illustrate some basic properties of the catalog, including correlations between the morphological and spectral properties of our large sample of cluster galaxies. A direct comparison of the spectral properties of the high redshift cluster and field populations suggest that the phenomenon of strong Balmer lines in otherwise passive galaxies (commonly called E+A, but renamed here as the k+a class) shows an order-of-magnitude increase in the rich cluster environment, compared to a more modest increase in the field population. This suggests that the process or processes involved in producing k+a galaxies are either substantially more effective in the cluster environment or that this environment prolongs the visibility of this phase. A more detailed analysis and modeling of these data will be presented in Poggianti et al. (1998).
astro-ph
astro-ph
Submitted: 1998 June 12; accepted: 1998 December 15 Preprint typeset using LATEX style emulateap j 9 9 9 1 n a J 0 2 1 v 3 6 2 1 0 9 9 / h p - o r t s a : v i X r a A SPECTROSCOPIC CATALOG OF 10 DISTANT RICH CLUSTERS OF GALAXIES Alan Dressler,1 Ian Smail,2 7 Harvey Butcher,4 8 Bianca M. Poggianti,3 , , Warrick J. Couch,5 Richard S. Ellis3 & Augustus Oemler Jr.1 1) The Observatories of the Carnegie Institution of Washington, 813 Santa Barbara St., Pasadena, CA 91101-1292 2) Department of Physics, University of Durham, South Rd, Durham DH1 3LE, UK 3) Institute of Astronomy, Madingley Rd, Cambridge CB3 OHA, UK 4) NFRA, PO Box 2, NL-7990, AA Dwingeloo, The Netherlands 5) School of Physics, University of New South Wales, Sydney 2052, Australia 6) Royal Greenwich Observatory, Madingley Rd, Cambridge CB3 0EZ, UK 7) Osservatorio Astronomico di Padova, vicolo dell’Osservatorio 5, 35122 Padova, Italy 6 , Submitted: 1998 June 12; accepted: 1998 December 15 ABSTRACT We present spectroscopic observations of galaxies in the fields of 10 distant clusters for which we have previously presented deep imaging with WFPC2 on board the Hubble Space Telescope. The clusters span the redshift range z = 0.37–0.56 and are the sub ject of a detailed ground- and space-based study to investigate the evolution of galaxies as a function of environment and epoch. The data presented here include positions, photometry, redshifts, spectral line strengths and classifications for 657 galaxies in the fields of the 10 clusters. The catalog comprises 424 cluster members across the 10 clusters and 233 field galaxies, with detailed morphological information from our WFPC2 images for 204 of the cluster galaxies and 71 in the field. We illustrate some basic properties of the catalog, including correlations between the morphological and spectral properties of our large sample of cluster galaxies. A direct comparison of the spectral properties of the high redshift cluster and field populations suggest that the phenomenon of strong Balmer lines in otherwise passive galaxies (commonly called E+A, but renamed here as the k+a class) shows an order-of-magnitude increase in the rich cluster environment, compared to a more modest increase in the field population. This suggests that the process or processes involved in producing k+a galaxies are either substantially more effective in the cluster environment or that this environment prolongs the visibility of this phase. A more detailed analysis and modeling of these data will be presented in Poggianti et al. (1998). Subject headings: galaxies: clusters: general – galaxies: evolution 1. INTRODUCTION The change with redshift observed in the proportion of star-forming galaxies in the cores of rich clusters was un- covered over twenty years ago, by Butcher & Oemler (BO, 1978, 1984), but it remains one of the clearest and most striking examples of galaxy evolution. Considerable ef- fort has gone into acquiring photometric information that would elucidate the physical processes active in distant clusters and their effects on the evolution of both the star-forming (Lavery & Henry 1994; Lubin 1996; Rakos & Schombert 1995; Rakos, Odell & Schombert 1997) and passive galaxies (Arag´on-Salamanca et al. 1993; Stanford, Eisenhardt & Dickinson 1995, 1998; Smail et al. 1998). Further impetus has been provided by observations of the recent transformation of the S0 population of clus- ters (Dressler et al. 1997), which may allow a closer con- nection to be drawn between the galaxy populations of distant clusters and the evolutionary signatures found in their local Universe counterparts (Caldwell & Rose 1997; Bothun & Gregg 1990). However, it was the advent of spectroscopic surveys of the distant cluster populations (e.g. Dressler & Gunn 1983, 1992, DG92; Couch & Sharples 1987, CS87; Barger et al. 1996; Abraham et al. 1996; Fisher et al. 1998) which un- covered the real breadth of the changes in galaxies in these environments, including several spectral signatures of evo- lutionary change, such as evidence for a strong decline in the star-formation rates of many cluster galaxies in the recent past. The advent of high spatial resolution imaging with the Hubble Space Telescope (HST ) provided a fur- ther breakthrough, giving morphological information on the galaxies in these distant clusters. This could be used to link the evolution of stellar populations in the galaxies with the evolution of their structure, in order to under- stand how the various galaxy types we see in the local uni- verse came to be. Pre- and Post-refurbishment HST obser- vations by two groups (Couch et al. 1994, 1998; Dressler et al. 1994; Oemler et al. 1997) were used in early attempts to correlate spectral evolution with morphological/structural data, and to provide some insight into the mechanisms that might be driving the strong evolution in the cluster galaxy population. These two programs were extended from Cycle-4 into the “MORPHS” pro ject, which accu- mulated post-refurbishment WFPC2 images for 11 fields in 10 clusters at z = 0.37–0.56, viewed at a time some 2– 4 h−1 billion years before the present day.9 The photomet- ric and morphological galaxy catalogs from these images were presented in Smail et al. (1997b, S97), while the data have also been used to study the evolution of the early-type galaxies within the clusters, using both color (Ellis et al. 1997) and structural information (Barger et al. 1998), the 8Visiting Research Associate at the Carnegie Observatories. 9We use q◦ = 0.5 and h = 1, where h = H◦ /100 km s−1 Mpc−1 . For this geometry 1 arcsec is equivalent to 3.09 h−1 kpc for our lowest redshift cluster and 3.76 h−1 kpc for the most distant. 1 2 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies Table 1 Telescope and Instrument Log Telescope Instrument Palomar 5.1-m WHT 4.2-m NTT 3.5-m COSMIC LDSS-2 EMMI λ (A) 3500–9800 3500–8300 3600–7800 Spectral Scale (A/px) 3.1 5.3 2.3 Spatial Scale # Nights (arcsec/px) Reference 0.40 0.59 0.27 19 6 2 Kells et al. (1998) Allington-Smith et al. (1994) Zijlstra et al. (1996) evolution of the morphology-density relation of the clus- ters (Dressler et al. 1997) and the masses of the clusters from weak lensing (Smail et al. 1997a). The aim of this paper is to combine the morphological information available from our HST images with detailed star-formation properties of the cluster galaxies derived from targeted spectroscopic observations. To this end we have used over 27 clear, dark nights over the past 4 years on the Palomar 5.1-m (P200),10 4.2-m William Herschel Telescope (WHT)11 and the 3.5-m New Technology Tele- scope (NTT)12 to assemble a large catalog of spectroscopic data on galaxies in these clusters. We combine these new observations with previously published spectroscopy from DG92 and present spectroscopic observations of a total of 424 cluster members, of which 204 have morphologies from our HST imaging, as well as 233 field galaxies (71 with HST morphologies). In addition, we have analyzed all of the spectra to provide equivalent width measurements on a uniform system for the entire sample. The spectral cata- logs, including line strength and color information, as well as the reduced spectra themselves in FITS format, are available at the AAS web site. A more detailed analysis of the spectroscopic data presented here will be given in Poggianti et al. (1998, P98). A plan of the paper follows. We start by discussing the observations and their reduction in §2. In §3 we then give the details of the redshift measurements, as well as our analysis to quantify the strengths of spectral features and information about our spectral classification scheme based upon these. We then present the spectral properties of galaxies in the catalog and relate these to the morpholo- gies of the galaxies from our HST images in §4, before discussing our results in §5. Finally in §6 we list the main conclusions of this work. 2. OBSERVATIONS AND REDUCTION ter members as it produces samples dominated by pas- sive spheroidal cluster members. We chose instead to base our ob ject selection upon galaxy morphology within the region covered by our WFPC2 imaging, while being ap- proximately magnitude-limited outside that area (selected from ground-based r or i CCD material to limits of r ∼ 22 and i ∼ 21). We note at this point that two of the cluster fields, A 370 Field 2 and Cl 0939+47 Field 2, lie outside of the central regions of their respective clusters (although we do also have observations of the core regions as well). The difference in the galaxy density between the fields should be kept in mind in the following analysis, although we will highlight such selection effects for individual figures when they are discussed below. Modelling of the sample selec- tion for the entire spectroscopic catalog is dealt with in more detail in P98. 2.2. Spectroscopic Observations The spectroscopic observations discussed in this paper were undertaken with a variety of facilities over the period 1993–1997. We list the instruments and telescopes em- ployed and the total number of nights used in Table 1. The basic details of the 10 clusters targeted in this study are listed in Table 2, this includes the mean cluster redshift, the one dimensional velocity dispersion (σcl , see §3.2), the redshift range used to define cluster membership (∆z ), the field center and the HST WFPC2 filters used in the ob- servations. The new spectra presented here are typically of high quality due to both the long exposure times em- ployed in our observations and the combination of the high efficiency of the multi-ob ject spectrographs and the large aperture of the telescopes used. We give in Table 3 the logs of the observing runs for the various telescopes. We list the mask identification, the dates of the observations, the total exposure time and the number of ob jects extracted from each mask (N). The slit width typically used was 1.5 arcsec, with slits between 10–20 arcsec long. The ex- act size of the region on the slit used to extract the galaxy The new spectroscopic observations discussed here were spectrum depended upon the relative signal to noise of the targeted at determining the membership of the numer- galaxy spectrum, but varied between 1.1–8.4 arcsec for the ous distorted and irregular galaxies revealed by our HST COSMIC spectra with a mean length of 3.9 ± 1.2 arcsec. WFPC2 images of the clusters, as well as gaining a more At the median redshift of the clusters in our catalog, the complete understanding of the star-formation properties of spectra thus sample a physical scale of ∼ (5 × 13)h−1 kpc. the general cluster population. With these aims, the ob- The exact details of the extraction and reduction of the ject selection is closer to that employed by DG92, than the spectra depends upon the instrument and set-up used. magnitude-limited selection criteria of CS87 and Barger However, the basic steps were the same for all the data et al. (1996). The latter approach has some claim to and we outline the procedures used for both the COSMIC making the subsequent analysis simpler, especially when and WHT/NTT data. The raw frames were debiased using the sample is selected in the near-IR. However, it is a the over-scan regions on the chip, before being trimmed. A very inefficient method for studying the faint, blue clus- 10The Hale 5-m of the Palomar Observatory is owned and operated by the California Institute of Technology. 11The William Herschel Telescope of the Observatorio del Roques de los Muchachos, La Palma, is operated by the Royal Greenwich Obser- vatory on behalf of the UK Particle Physics and Astronomy Council. 12Based in part on observations collected at the European Southern Observatory, La Silla, Chile. 2.1. Selection of Spectroscopic Targets Dressler et al. Table 2 Properties of The Clusters 3 Cluster z A 370 Cl 1447+26 Cl 0024+16 Cl 0939+47 Cl 0303+17 3C 295 Cl 0412−65 Cl 1601+42 Cl 0016+16 Cl 0054−27 0.3741 0.3762 0.3928 0.4060 0.4184 0.4593 0.5074 0.5388 0.5459 0.5608 σcl (km s−1 ) 1170 1470 1150 1260 1310 1630 700 1210 1660 1180 ∆z 0.3589–0.3873 0.3621–0.3857 0.3755–0.4081 0.3879–0.4173 0.4018–0.4338 0.4464–0.4733 0.5024–0.5130 0.5100–0.5473 0.5300–0.5601 0.5520–0.5770 R.A. (J2000) Dec. (J2000) 02 39 52.6 −01 34 18 14 49 29.3 +26 07 52 00 26 35.7 +17 09 46 09 42 56.1 +46 59 12 03 06 12.9 +17 20 08 14 11 10.5 +52 12 11 04 12 50.1 −65 50 44 16 03 12.0 +42 45 26 00 18 33.3 +16 26 16 00 56 59.0 −27 40 20 HST Filters Comment F555W/F814W, F702W Center (F702W), outer field (F555W/F814W) F702W F450W/F814W F555W/F814W, F702W Center (F702W), outer field (F555W/F814W) F702W F702W F555W/F814W F702W F555W/F814W F555W/F814W Spectra from DG92 two dimensional flatfield was constructed by dividing the flatfield exposure by a low-order fit in the dispersion direc- tion. The data frame was then divided by this normalized flatfield, this served to correct for the pixel-to-pixel re- sponse of the detector. The sequence of data frames for each mask taken on a single night were then checked for spatial offsets between the exposures arising from flexure ∼ 0.2 pixels in the spectrograph (these are typically only < for COSMIC in the course of a night). If necessary the exposures were shifted in the spatial and/or dispersion di- rection to align them and then combined with a cosmic- ray rejection algorithm using the IRAF task imcombine. This produced a two dimensional image of the mask ex- posure clean of cosmic ray events. These frames were then geometrically remapped to align the spectra along the rows of the detector. This step is necessary to remove the distortion of the spectra on the detector introduced by the spectrograph optics. The distortion is only a large effect for ob jects in slits near the edge of COSMIC’s large 13.7′ × 13.7′ field of view, although aligning the spectra also helps when tracing some of the faintest ob jects. The distortion of the spectra are mapped using the positions of the emission lines in the arc exposure taken after every science exposure. The positions of ob jects in each slit on the remapped frame, as well as regions of clear sky sur- rounding them, were then defined interactively using the IRAF package apextract. The exact position of the ob- ject within the slit was traced in the dispersion direction and fitted with a low-order polynomial to allow for atmo- spheric refraction. The spectra were then sky-subtracted and extracted using optimal weighting to produce one di- mensional spectra. The arc exposures associated with each science exposure were remapped and extracted in exactly the same manner (although with no sky-subtraction) and these were used to determine the wavelength calibration for the science exposure. We estimate our wavelength scale is good to 0.2A rms. Finally, the one dimensional spectra were smoothed to the instrumental resolution, ∼ 8A, and rebinned to 10A per pixel to make them more manage- able. The spectra obtained with COSMIC have not been flux calibrated. The WHT and NTT spectra have been reduced using the LEXT package, purposely written for reducing LDSS– 2 spectra, and the MIDAS software package. What fol- lows is a brief description of the reduction procedure gen- erally adopted. A number of twilight and dome flatfields, and several arc frames were obtained for each mask, as well as numerous bias frames and long–slit spectra of standard stars for flux calibration (at least one star per night). The raw frames were first debiased and then divided by the corresponding normalized flatfield. They were then cali- brated in wavelength with the arcs frames obtained either with a CuAr or HeArNe combination of lamps. The sky– subtraction step was performed with an interactive choice of the spatial limits of the spectrum, which was then ex- tracted summing the counts weighted with a Gaussian. The long–slit stellar spectrum was reduced in a similar way as the target spectra and a response function was de- rived by the comparison with a tabulated spectrum. Each spectrum was flux–calibrated in Fν by dividing for this re- sponse function. In the case of the WHT and NTT spectra each exposure of a given mask was reduced and calibrated separately, before all the spectra of a given galaxy were coadded; no smoothing or rebinning was applied. The full digital catalog of FITS spectra collected for this program is distributed in electronic form on the AAS web site. These spectra are also available from: http://www.ociw.edu/∼irs. 3. SPECTROSCOPIC ANALYSIS The full catalog of ob jects observed spectroscopically in the 10 clusters is given in Tables 4 (the complete tables are included on the AAS web site as well as being available from http://www.ociw.edu/∼irs). This has been split into “Cluster” and “Field” samples as described below. The ta- bles list not only the spectral information on the galaxies, but also any available morphological and photometric data from S97 and DG92. A key to the various parameters and the format of the tables are given in Table 5. We now describe in more detail the measurement of some of the spectral parameters listed in Tables 4. 3.1. Spectral Measurements The quality of the spectra, both in terms of signal-to- noise and sky-subtraction, was visually assessed by AD for al l of the spectra presented. The spectra are graded on a 4–point range, with q = 1 signifying the best and q = 4 the worst quality. Of the complete catalog 17% have q = 1, 47% with q ≤ 2 and 89% are q ≤ 3. Spec- tra with q ≤ 3 have sufficient signal to noise (S/N) for not only measurement of a redshift, but also to quantify the strength of any spectral features present. From the continuum regions around the [Oii]λ3727 and Hδ lines we estimate median S/N of 40.2 (q = 1), 28.3 (q = 2) and 19.7 (q = 3), with lower limits to the S/N of 20.9, 10.6 and 4.6 4 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies Table 3 Log of New Spectroscopic Observations Comments Using ‘old’ COSMIC CCD Target/Mask Date Exposure time (ks) COSMIC/P200 Cl 0024+16 EW–1 Cl 0024+16 EW–2 Cl 0024+16 NS–1 Cl 0024+16 NS–2 A370 MS–1 A370 MS–2 A370 MS–3 Cl 0303+17 EW–1 Cl 0303+17 EW–2 Cl 0303+17 NS–1 Cl 0939+47 NS–B Cl 0939+47 EW–B Cl 0939+47 EW–F Cl 1601+42 CM–1 Cl 1601+42 CM–2 Cl 1601+42 CM–3 5–6 Aug 1994 29 Nov 1994, 1–3 Dec 1994 29–30 Oct 1994 19 Aug 1995, 25–26 Sept 1995 25–26 Sept 1995 27–28 Nov 1995 4 Oct 1997 29 Nov 1994, 1 Dec 1994 2 Dec 1994 3 Dec 1994 13 Dec 1993 29–30 Nov 1994, 1 Dec 1994 1–2 Dec 1994, 24 Feb 1995 28–29 May 1997 18–19 June 1996 11–12 May 1997, 29 May 1997 LDSS-2/WHT 24 Mar 1993 Cl 0939+47 WA–1 25 Mar 1993 Cl 0939+47 WA–2 26 Mar 1993 Cl 0939+47 WA–3 Cl 0939+47 WA–4 26 Mar 1993 Cl 0939+47 MA–A,MA–D 27–28 Apr 1995 29 Apr 1995 Cl 0939+47 MB–A,MB–D 3C 295 MA–A,MA–D 27 Apr 1995 Cl 1447+26 WA-1 Cl 1447+26 MA-A,MA-D Cl 1601+42 MA-A,MA-D 25-26 Mar 1993 28–29 Apr 1995 28 Apr 1995 Cl 0054−27 MA-1 Cl 0054−27 MA-2 Cl 0412−65 MA-1 Cl 0412−65 MA-2 23 Nov 1995 24 Nov 1995 23 Nov 1995 24 Nov 1995 EMMI/NTT 18.8 20.8 27.0 20.0 10.8 24.0 16.2 25.0 15.0 15.0 10.8 38.2 36.0 22.2 21.6 30.2 14.4 14.4 10.8 9.0 9.0 9.0 9.0 10.8 12.6 9.0 10.2 10.8 9.0 8.6 N 37 34 35 29 25 29 22 30 30 27 18 35 34 20 22 24 10 7 7 6 10 7 6 23 9 8 15 21 15 21 respectively for these three quality classes. Repeated ob- servations suggest that the redshifts of q = 1 and q = 2 cases are correct at a confidence of greater than 98%, and that q = 3 cases are correct at a confidence of greater than 90%. In contrast, those spectra with q = 4 are of sufficient S/N to provide only a redshift, which may be uncertain in a significant number of cases. Redshifts were measured from the spectra interactively using purpose-written software that compares the wave- lengths of redshifted absorption and emission lines with features in the spectra. Whenever possible we used a num- ber of features to estimate the redshift, and only in a very small number of cases is a redshift based on only a sin- gle feature — these instances are noted in the comments in Tables 4. We list in column 24 of Tables 4 the main features used to identify the galaxy redshifts. For concise- ness we have used the following abbreviations to identify the lines: Babs, Balmer absorption lines; Ha, Hα; Hb, Hβ ; Hd, Hδ ; He, Hǫ; Heta, Hη ; Hg, Hγ ; Hth, Hθ; Hz, Hζ ; G, G-band; H&K, Ca H or K; Mg, Mg-B; Na, Na-D; OII, [Oii]λ3727; OIII, [Oiii]λ4959,5007; bk, 4000A break; MgII, Mgiiλ2799; CIII, Ciii]λ1909; CIV, Civλ1549; FeI, Feiλ5268; NII, [Nii]λ6583; SII, [Sii]λ6716,6731. The strength of emission and absorption features in the spectra were measured using purpose-written software, al- lowing the positioning of the continuum to be defined in- teractively. We give the restframe equivalent widths (EW) for [Oii]λ3727 and Hδ in columns 5 and 6 of Table 4A and 4B, in all instances a line seen in emission is given a negative value and is quoted in A. The presence and strength of these lines is used in the spectral classification scheme discussed in §3.2. If other lines in the spectrum were measurable we list their EW in the comments. We give line strengths for not only those galaxies observed for this work, but also those from the early survey of DG92. The D4000 measurements have been similarly placed on a consistent system. These are measured using wavelength intervals as defined in Dressler and Shectman (1987). The COSMIC data shared a common relation of counts to flux, but were not flux calibrated per se. A multiplicative cor- rection of 1.34 to convert the measured D4000 to true D4000 for these data was derived by comparing the COS- MIC spectra of repeated ob jects with the equivalent flux calibrated DG92 spectra. This procedure, though imper- fect, generates reasonable and consistent results, as shown by multiple COSMIC observations of the same galaxies. We have a total of 31 repeat observations, both in- ternally within the datasets from a single telescope, and between telescopes. We find median rms scatters of σ(zCOSMIC − zDG92 ) = 0.0018 (N = 14), σ(zCOSMIC − zWHT ) = 0.0009 (N = 2) and σ(zCOSMIC − zCOSMIC ) = 0.0005 (N = 7) for those spectra with q ≤ 3, and no systematic offsets between any of the individual datasets: < zCOSMIC − zDG92 >= 0.0007, < zCOSMIC − zWHT >= −0.0009. We therefore conclude that there are no sig- nificant offsets between the redshifts from the different datasets and hence we are confident that we can include all the observed ob jects in our analysis. Finally, we quantified the detectability of [Oii] and Hδ in our spectra. This enabled us to derive the lower lim- Dressler et al. 5 Table 5 Notes on The Parameters in Tables 4 Column Parameter Units Format Comment 1..... 2..... 3..... 4..... 5..... 6..... 7..... 8..... 9..... 10..... 11..... 12..... 13..... 14..... 15..... 16..... 17..... 18..... 19..... 20..... 21..... 22..... 23..... 24..... 25..... CLUSTER ID z Q [Oii] Hδ D4000 CLASS δRA δDec IDHST X Y MORPH T D INT MAG COL MAGDG COLDG RUN MASK FEATURES COMMENTS A A arcsec arcsec Pixels Pixels Mag Mag Mag Mag A6 I4 F7.4 A1 I2 F7.1 A1 F4.1 A1 F5.2 A11 F7.1 F7.1 I5 I5 I5 A12 I2 I2 A6 F5.2 F5.2 F6.2 F6.2 A6 A10 A23 A130 Cluster ID in spectroscopic catalog for cluster Redshift Redshift quality – a “:” indicates questionable identification Quality of spectrum: 1=High, 4=Low. Restframe equivalent width of [Oii] 3727 Quality of [Oii] 3727 EW measurement (a “:” indicates questionable) Restframe equivalent width of Hδ, −ve indicates emission Quality of Hδ EW measurement (a “:” indicates questionable) Break strength index Spectral classification in scheme described in §3.3 RA offset from field center in Table 2 Dec offset from field center in Table 2 ID in photometric catalog for clustera X coordinate on WFPC2 framea Y coordinate on WFPC2 framea Galaxy morphologya T-typea Visual disturbance indexa Interpretation of disturbancea Total magnitude in F702W/F814W from WFPC2 framea,b Aperture color from WFPC2 framea,c Magnitude from ground-based imaging published in DG92d Color from ground-based imaging published in DG92d Code giving details of observing rune Mask and ob ject slit identifier Spectral features identified, see §3.1 Description of features in spectrum a See S97 for more details. b Magnitudes are in F702W for Cl 0303+17, Cl 0939+47, 3C 295, Cl 1447+26 and Cl 1601+42, and in F814W for Cl 0016+16, Cl 0024+16, Cl 0054−27, A 370 Field 2, Cl 0412−65, Cl 0939+47 Field 2. c WFPC2 V555 − I814 color information is available for: Cl 0016+16, Cl 0054−27, A 370 Field 2, Cl 0412−65, Cl 0939+47, and B450 − I814 colors for Cl 0024+16. d Aperture r-band magnitude from DG92, colors are aperture (g − r) measurements in all instances. e [P/W/N]<MONTH><YEAR>, P=Palomar 5-m, W=WHT, N=NTT, or DG92. its on the strength of these spectral features below which we would not have identified them. Achieving this aim was not straightforward because the code that best mea- sured the equivalent widths, which is based on a gaus- sian line-fitting program written by Paul Schechter, does not perform well when the lines are weak or undetectable. For this reason, when we measured the strengths of fea- tures in those galaxies where the feature was not clearly seen, we by necessity had to measure equivalent widths using the standard technique of obtaining the continuum level from straddling continuum bands, and measuring the decrement or increment in signal relative to the continuum in an interval containing the feature. We made such mea- surements of [Oii] and Hδ EW for all COSMIC spectra with qualities q ≤ 3 of cluster members in Cl 0939+47 and Cl 0024+16, a total of 79 galaxies. The intervals are, again, as defined in Dressler and Shectman (1987). For weak, but measurable, cases the line-fitting and flux-summing techniques give equivalent results, though for strong ab- sorption lines, in particular, the latter seems to underes- timate the strength of the feature, apparently by allowing the wings of the line to lower the continuum level. We believe, however, that the two scales for measuring equiv- alent widths are interchangeable for the purpose of looking for weak features. The results of these tests are shown in Fig. 1a and Fig. 1b, where we have plotted the equivalent widths as a function of signal-to-noise ratio in the continuum bands straddling the feature. In Fig. 1b we show that the galaxies that were designated by inspection as emission line types all have [Oii] EW stronger than −3A, while those desig- nated as having no emission lines (spectral types: k, k+a, or a+k, see §3.3) have [Oii] EW weaker than −4A. In fact, the latter are consistent with non-detections: for 37 non-emission line members, the median EW is +0.4A with quartiles of −1.0 to +2.6. There is only a weak trend with signal-to-noise ratio. We conclude from these data that we are complete for [Oii] stronger than −5A, with a high level of completeness down to −3A. In other words, even at the modest signal-to-noise ratios of these spectra, none of the galaxies classified as non-emission types are likely to have emission at greater than the −3A level, and certainly none have emission stronger than −5A (this limit corresponds to “absent” in Table 6). In Fig. 1b we show a similar diagram for the same sam- ple, this time for Hδ . Because it is weaker and in ab- sorption, Hδ is a more difficult feature to measure; this is apparent from the stronger trend with signal-to-noise ratio. However, as for [Oii], the separation of those galax- ies which are designated by inspection as having moderate Balmer line strengths (k+a, a+k, and e(a), see §3.3) from the non-Balmer galaxies (k and e(c) and e(b) types), is confirmed by the ob jective measurements. The boundary is around 2–3 A, below which point we are unable, except at high S/N > 50, to discern the difference between the presence or absence of Hδ . We conclude from these data that we are complete above equivalent widths of +5A, and mostly complete above +3A. It is worth commenting that some of the points with large negative equivalent widths for Hδ arise from strange continuum levels, rather than from the feature seen in emission (although there is at least one clear case of Hδ in emission, a rare phenomena among luminous galaxies). 6 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies k k+a,a+k e(c),e(a), e(b) -40 -20 0 0 10 20 30 40 50 15 10 5 0 -5 0 k k+a,a+k e(c),e(b) e(a) 20 40 60 80 100 Fig. 1. a) Measured [Oii] equivalent widths versus the S/N in the straddling continuum. The data are for 79 cluster members with q ≤ 3 in the clusters Cl 0939+47 and Cl 0024+16. There is a clean separation at ∼ −4A between the types for which [Oii] was found by inspection and those for which it was judged to be absent. b) The equivalent plot for the measured Hδ equivalent widths. Hδ strengths greater than 3A are clear detections, according to the distribution, with 2–3A strengths ambiguous, particularly at low S/N. 3.2. Cluster Membership As was noted above, Table 4 is split into two parts on the basis of whether a galaxy is classed as a “Cluster” member or “Field”. To accomplish this we define redshift ranges for the various clusters; these ranges are purposefully cho- sen to be large to ensure that we retain any galaxies in the large-scale structure surrounding the clusters, while at the same time minimizing the contamination by field galax- ies. In Fig. 2 we show the redshift distributions for the individual cluster fields; in each panel the inset provides a more detailed view of the velocity distribution close to the cluster mean. The bin size in these plots has been ar- bitrarily chosen and may artificially enhance or suppress the visibility of any structures within the clusters. We list the resulting mean redshift, restframe velocity dispersion and redshift range defining each cluster in Table 2. We reiterate that the velocity dispersions are likely to be over- estimates of the true dispersion of the well-mixed cluster population. We also list in Table 2 the number of member galaxies in our catalog for each cluster. Using these defini- tions our catalog contains a total of 424 cluster members and 233 field galaxies. The redshift distribution for all galaxies classed as field is shown as the open histogram in Fig. 3; the galaxies with HST morphologies are shown as the filled histogram. The median redshift of the whole field sample is < z >= 0.42, while for the morphological sub-sample it is slightly higher at < z >= 0.46 (Fig. 3). These values are very similar to the median redshift of our 10 clusters, < z >= 0.44, allow- ing us to easily compare the broad properties of the cluster and field samples. A total of 20 stars were observed (all in either the flanking fields or from the earlier DG92 ob- servations); these are included at the bottom of Table 4b, but we do not discuss them further. In Fig. 3 we may be seeing some evidence for a deficit in the total field redshift distribution, between z ∼ 0.4–0.6, which would result from the inclusion of a few field galax- ies in the cluster catalog. This would include galaxies in the supercluster environment, if any, in which the clusters reside, or truly unassociated galaxies relatively far from the cluster but within the wide velocity limits imposed by the cluster’s velocity dispersion. To estimate the extent of this effect we use two approaches. Firstly, a conservative upper limit on the deficit comes from linearally extrapo- lating the trends of number versus redshift in the field at z < 0.35 and z > 0.60 to limit the likely number of field galaxies in the intervening redshift range. From this we ∼ 160 field galaxies in the estimate that there should be < range z = 0.3–0.6, compared to the observed number of 92, giving an upper limit on the deficit of ∼ 70 galaxies, ∼ 7 per cluster. Alternatively, using the regions where or < the redshift limits of the cluster and field samples overlap between different clusters we estimate the contamination from random, unrelated field galaxies is of the order of 1.0 ± 0.7 galaxies per cluster in our largest velocity range. We conclude therefore that the contamination from galax- ies unrelated to the cluster, or its supercluster, does not exceed 7 galaxies per cluster and is probably closer to 1–2 galaxies. 3.3. Spectral Classification To assess the distribution in the star-formation prop- erties of galaxies in our catalog we have found it useful to classify their spectra into a number of classes. These classes are broadly based upon those used by DG92 and CS87, however, the number of classes has been expanded to better cover the full range of features seen in our large sample. We have also used the properties of low red- shift integrated spectra (Kennicutt 1992) and the expected characteristics from spectral modelling to help us define the limits of some of the classes. In revising the classifi- cations we therefore found it necessary to redefine some of the boundaries previously used for the spectral classes. Hence, to reduce confusion between our new classes and Dressler et al. 7 Fig. 2. Redshift distributions for the fields of our 10 clusters. We show the redshift range z = 0–1 in the full plot and then, in the insets, an expanded region (width ∆z = 0.08) centered on the cluster redshift, in the inset panels we show the cluster members as a filled histogram and the field galaxies as open. The redshift axis in the inset panel is marked with ∆z = 0.01 increments and the vertical axis is the same as the main panel. therefore have detailed morphological information (71 galaxies). those used previously we adopt a new nomenclature and give this and the details of the classification scheme in Ta- ble 6. We show a schematic representation of this spectral classification in Fig. 4. It should be noted that for those spectra where sky residuals or the available spectral range precluded the observation of one of the diagnostic spec- tral features, we have made used the strength of the other Balmer series lines (if Hδ was unobservable) or emission lines (if [Oii] was unobservable) to identify the most likely spectral class. In the few cases where this has been done comments are included in Table 4. Fig. 4. A schematic representation of the spectral classification scheme used in this work. We show the regions of the Hδ–[Oii] equiva- lent width plane populated by the various spectral types. Those spectral classes not based upon the line strengths of Hδ and [Oii] (e.g. CSB, e(n), etc.) are not marked. Briefly the overlap between the new system and pre- vious ones can be summarized as follows: we retain the general features of the DG92 system, including k-type and the general class of “e” (emission) galaxies. However, we replace the mixed nomenclature “E+A” with k+a (follow- ing the suggestion of Franx 1993) and a+k, depending on the strength of the Hδ Balmer line. We also subdivide emission line galaxies into e(a) types (with strong Balmer absorption), e(c) for those with weak or moderate Balmer absorption, and e(b) for those with very strong [Oii] (this can sometimes be combined with e(a) for galaxies with both strong [Oii] emission and strong Balmer absorption). This nomenclature reflects the nature of the spectra, with Fig. 3. Redshift distribution for galaxies classed as non-members in the fields of the 10 clusters. The open histogram gives the total redshift distribution for the field galaxies (233 galaxies), the filled histogram is those field galaxies which lie within the WFPC2 field and for which we 8 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies e(a) indicating a population of A stars, e(b) a spectrum similar to that expected for a burst of star-formation and e(c) a spectrum for a system undergoing a more constant SFR. In comparison to other earlier work, the PSG and the HDS galaxies of CS87 fall mostly into the a+k and k+a classes. The CS87 “Spiral” types are placed in e(c) and e(a); however, the SB galaxies are not the same as our type e(b), because the criteria for these in CS87 was not based on [Oii] strength. We note that spectral classes described in Table 6 can be grouped in three main categories: pas- sive (k); past star-forming (k+a and a+k) and currently star-forming (e(c), e(a) and e(b)). AGN spectra (e(n)) are excluded in this division (Table 6). In Column 8 of Table 4 we include a photometric clas- sification in the case of the bluest galaxies. These are labeled “Color Starburst” (CSB) if their restframe color is bluer than that expected for a low metallicity model galaxy with an increasing star-formation rate (P98). This allows us to conservatively identify those galaxies whose very blue colors can only be explained with a current star- burst, whatever their spectral type may be. Fig. 5. Representative spectra from each of the spectral classes in our adopted scheme (Table 6, Fig. 4). These are plotted with arbitrary ver- tical scaling and in the restframe. The galaxies are all cluster members with q = 1 and come from Cl 0939+47 and Cl 0024+16. The spectra are not fluxed. To better illustrate the properties of the new classifica- tion scheme we show in Fig. 5 a high-quality, representa- tive member of each class from our catalog. In Table 7 we give the distribution of spectral classes within the different clusters (for q ≤ 4), as well as the total numbers across all the clusters and the equivalent values for our field samples. As can be seen, the clusters are populated by a wide variety of spectral classes, although comparisons between clusters are not simple owing to the different apparent magnitudes of the samples and the attending variation in the typical quality of the spectra. Table 7 also lists the equivalent numbers of galaxies in each spectral class for which we have morphological information. 4. BASIC PROPERTIES AND CORRELATIONS OF THE DATA To start the discussion of the spectroscopic sample we have assembled, we review the basic properties of the sam- ple as a whole. We focus on a few of the correlations be- tween the various properties of the galaxies in the sample, in particular the relationships between the morphological, spectral, and kinematic characteristics of certain classes of cluster galaxies. In the following discussion we will in- clude the uncertain spectral classes (marked with a “:” in Tables 4), unless otherwise stated. 4.1. Luminosity Functions for the Morphological Classes In order to draw conclusions from our spectroscopic study in the context of the broader morphological cata- log (S97), we need to compare the sampling in absolute magnitude of the two catalogs. Fig. 6a shows the abso- lute magnitude distribution for galaxies in the spectro- scopic catalog for which ground-based r-band photometry is available. This filter approximates V in the restframe for all 10 clusters. Our assumption of a single K-correction (from an spectral energy distribution (SED) corresponding to a present day Sbc) introduces only small errors into the ∼ 0.06 mags for E/S0 and Sd/Irr magnitude distribution (< SED). Fitting a Schechter function to the bright-end of the distribution in Fig. 6a, we obtain a characteristic magni- tude of M ∗ V = (−20.64 ± 0.16) + 5 log10 h (for a fixed faint- end slope of α = −1.25 as adopted in S97). This is to be compared to a fit obtained to the morphological counts in the cluster fields corrected for likely contamination in the manner described in S97. Fitting to the composite luminosity function of all morphological types across the 10 clusters we find M ∗ V = (−20.79 ± 0.02) + 5 log10 h (for α = −1.25). This good agreement indicates that the spec- troscopic catalog fairly samples the morphological catalog for MV < −19 + 5 log10 h. Fig. 6b shows how the spectroscopic sampling compares as a function of morphological type within the clusters. This is achieved by comparing the spectroscopic sample for MV < −19 + 5 log10 h to the field-corrected morpho- logical counts of S97. There is no significant trend with morphological type except for the selection effect, dis- cussed in §2.1, built into the original sample selection: the Sd/Irregular galaxies are oversampled relative to the E–Sc types (although there is considerable uncertainty in the statistical correction for field galaxies in this bin, S97). This plot allows us to quantify and correct for the sample selection in our analysis as required. 4.2. Luminosity Functions for the Spectral Classes The absolute magnitude distribution of the spectral classes defined in this paper will be important to un- derstanding their relationships within the framework of galaxy evolution models. Fig. 7a shows that the magni- tude distributions brighter than MV = −19 + 5 log10 h for spectral classes k, k+a and a+k are statistically indistin- guishable. In contrast, the e(a) and e(c) classes appear to systematically fainter than the k class; this difference is ∼ 95% confidence limit using two-sample confirmed at the > Kolmogorov-Smirnoff tests. It is important to keep in mind the “completeness” limit of the spectroscopic catalog Dressler et al. 9 Class k k+a a+k e(c) e(a) e(b) e(n) e ? CSB EW [Oii] 3727 (A) absent absent absent yes,< 40 yes ≥ 40 ... yes ? ... EW Hδ (A) < 3 3–8 ≥ 8 < 4 ≥ 4 ... ... ? ? ... Table 6 Spectral Classification Scheme Color Comments ... ... ... ... ... ... ... ... ... very blue passive moderate Balmer absorption without emission strong Balmer absorption without emission moderate Balmer absorption plus emission, spiral-like strong Balmer absoprtion plus emission starburst AGN from broad lines or [Oiii] 5007/Hβ ratio with at least one emission line but S/N too low to classify unclassifiable photometrically-defined starburst Table 7 Spectral Samples Cluster Ntot k k+a e(c) e(b) e(n) e ? a+k e(a) Full Sample 0 37 3 0 6 0 13 2 7 6 6 4 1 2 0 0 4 0 3 5 0 0 18 44 7 1 1 12 13 4 6 1 15 6 1 60 Morphological Sample 7 0 5 2 0 0 2 0 0 4 0 6 5 2 6 2 3 4 2 0 5 0 0 0 0 0 5 2 5 4 0 0 0 10 28 21 Field A 370 Cl 1447+26 Cl 0024+16 Cl 0939+47 Cl 0303+17 3C 295 Cl 0412−65 Cl 1601+42 Cl 0016+16 Cl 0054−27 Total Field A 370 Cl 1447+26 Cl 0024+16 Cl 0939+47 Cl 0303+17 3C 295 Cl 0412−65 Cl 1601+42 Cl 0016+16 Cl 0054−27 Total 233 40 21 107 71 51 25 10 58 29 12 424 71 14 10 42 31 28 20 2 28 22 7 204 36 26 7 47 31 14 10 0 33 13 5 186 11 9 4 25 16 9 7 0 20 10 4 104 74 8 6 21 10 12 2 0 3 1 1 64 25 1 3 3 2 5 2 0 1 1 0 18 39 1 1 6 2 6 0 1 2 1 0 20 11 1 1 3 0 4 0 0 1 0 0 10 3 0 0 2 1 0 3 0 1 0 0 7 0 0 0 0 0 0 3 0 1 0 0 4 25 0 0 2 1 5 1 1 0 0 3 13 9 0 0 1 0 1 1 0 0 0 1 4 12 1 0 2 0 0 0 7 0 0 2 12 3 1 0 0 0 0 0 2 0 0 2 5 10 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies estimated in §4.1, which means that these differences could be larger, and, for example, the apparent peak in the lumi- nosity distribution of e(a)’s in Fig. 7a may be partly an ar- tifact of incomplete sampling. The difference between the k class and the fainter e(b) class is clearly significant: the likelihood that the two samples are drawn from the same luminosity distribution is only log10 P ∼ −4.6. Again, the difference may be larger still, owing to the incomplete sampling below (MV = −19 + 5 log10 h). We know of no selection effect in our study that would cause us to miss bright e(b) cluster galaxies. As we discuss in P98, the fact that the galaxies that we have identified as bursting are fainter than the other classes is significant, and discourag- ing for models that attempt to interpret these starbursts as progenitors for galaxies with strong Balmer lines in their spectra. The cluster sample defined by our redshift measure- ments also allow us to unambiguously derive, for the first time, the absolute magnitude distributions as a function of morphology, again for MV ≤ −19 + 5 log10 h. Fig. 7b shows a broad similarity between the absolute magnitudes of early- and mid-type disk systems (S0–Sa–Sb–Sc). Com- pared to these, elliptical galaxies show a systematically brighter distribution, and irregular galaxies exhibit a tail of fainter systems. These trends are in good agreement with what is seen in low-redshift clusters. 4.3. Morphological Properties of the Cluster Galaxies Fig. 10. A comparison of the distribution of morphological type within each spectroscopic class, for both cluster and field galaxies. We briefly discuss evidence for interactions and mergers on the spectral classes of galaxies in our cluster samples. (We also comment on this issue in §5.2 which deals with the kinematics of the different cluster populations.) We show in Fig. 9a the distribution of disturbance class within the different spectral classes. The image disturbance, D, is a visual classification of the degree to which the galaxy’s structure appears distorted or disturbed (S97) compared to a typical low-redshift galaxy of the same morphologi- cal type. The D class correlates well with the asymmetry of the galaxy’s light profile (S97). Fig. 9a suggests that the spectral properties of the galaxies broadly correlate with the degree of image distortion and disturbance, the active and recently active populations having more galax- ies classed as strongly asymmetric or distorted. However, looking at Fig. 9b we see an arguably stronger correlation between morphology and D with a pronounced shift to- wards higher D values in going to later-types (Sb–Sd/Irr). This could be due to a failure on our part to actually separate disturbance from a natural trend toward more irregular morphology for late-type systems, but the large number of D ≥ 2 Sc galaxies (a type that is generally symmetric for low-redshift galaxies) suggests that the ef- fect is real.13 . If so, it most likely reflects the greater fragility of disks (compared to bulges) to perturbations, and the greater frequency of perturbations at higher red- shift. However, we see that this effect does not appear to be result of the high density cluster enviroment: Fig. 9b shows that 50 ± 8% of the cluster Sb–Sc–Sd/Irr galaxies have D ≥ 2, a proportion similar to that seen in the late- type field population, 60 ± 11%. The same effect is seen at low-redshift (Hashimoto & Oemler 1999). What do the galaxies in our spectral classes look like? We illustrate the morphologies of the cluster members within each spectral class in Figs. 8. The general trend towards later-types in the active spectral classes is clear. The passive spectral classes are dominated by early-type galaxies, particularly ellipticals. The correspondence of morphology and spectral properties, the same as found for low-redshift analogs. indicates that a substantial fraction of the luminous ellipticals of these clusters was in place by z ∼ 0.5 (Ellis et al. 1997, Dressler et al. 1997). The e(c) spectra are generally associated with disk galaxies, most of them familiar spirals and irregulars. This is true of some of the e(a)’s as well, but this class also includes many disk systems that look more disturbed than typical present-day spirals. The k+a/a+k class does include some elliptical galaxies, but the ma jority are disk galaxies, a few of which have an irregular or disturbed appearance. The significance of the correlation of morphology and spectral class are discussed further in B98. 13This tendancy of intermediate-redshift disk galaxies to appear more asymetric than low-redshift galaxies of similar type has been reported in essentially all studies of this type 4.4. Spectroscopic Properties of the Cluster Galaxies Dressler et al. 11 Fig. 6. a) The absolute magnitude distribution, in the restframe V -band, for the spectroscopic sample of cluster (solid histogram) and field (dashed histogram) galaxies (for those galaxies with DG92 photometry). The solid curve is the best-fit Schechter function to the cluster members, using a fixed faint-end slope of α = −1.25. The characteristic luminosity derived from the fit is M ∗ V = −20.64 ± 0.16 + 5 log10 h. The dotted line shows the fit to the luminosity functions derived from the morphological counts in the frames, corrected for field contamination (S97). This fit is shown for α = −1.25 and with arbitrary vertical scaling. The good agreement of the two distributions shows that the spectroscopic catalog provides a representative luminosity distribution in the clusters at MV < −19 + 5 log10 h. b) The numbers of the different morphological types in the spectroscopic catalog (hatched histogram) brighter than MV = −19 + 5 log10 h. The filled histogram indicates the total numbers expected from the observed morphological counts in the clusters to the same depth, after correcting for field contamination (see S97). Fig. 7. a) The absolute magnitude distribution in the rest-frame V -band for cluster members separated into the different spectroscopic classes for both the HST fields and the full sample. In this panel we convert the ground-based r-band photometry to rest-frame V -band. b) The same distribution, but now separated on morphological type and using the HST photometry. The K-correction applied in both cases assumes a spectral energy distribution (SED) similar to a local Sbc galaxy, this introduce a typical systematic error of only ∼ 0.06 mags for E/S0 and Sd/Irr SEDs. 12 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies Fig. 9. a) Histograms indicating the distribution of spectral types within the different disturbance classes, D, for both cluster and field populations (hatched and filled histograms respectively). D ≥ 2 denotes strongly a asymmetric or disturb light distribution. b) The numbers of galaxies in the different disturbance classes as a function of galaxy morphology. Note the broad similarity of the cluster and field distributions for the later-type spiral galaxies. In Fig. 10 we quantify the distribution of morphologi- cal type for the various spectral classes, for both cluster members and field galaxies. The strong, though broad, relation between morphology and star-formation seen in low-redshift galaxies is present in this intermediate red- shift sample as well. Looking at the star-forming popula- tion which causes the Butcher-Oemler effect we see a clear tendency for these galaxies to be predominantly late-type systems (Couch et al. 1994, 1998; Dressler et al. 1994; Oemler et al. 1997), although here there is a tail of earlier- types (at least in the e(a) and e(c) classes). These ac- tive early-type (E and S0) galaxies comprise a higher frac- tion of the field population than they do in the clusters. The two “recently-active” classes, k+a and a+k, appear to have morphological distributions which are intermedi- ate between the passive and active cluster populations. There seems to be a clear distinction between k+a and a+k in the sense that the latter are of later morphologi- cal type, though the small number of a+k types limits the statistical certainty of this result. Fig. 11. The cumulative distribution of [Oii] 3727 EW for three inde- pendent morphological bins for both cluster (solid line) and field (dotted line) populations. It is interesting that, although the passive cluster popu- lation is dominated by elliptical and S0 galaxies, there is a significant number of later types, stretching out to Sd/Irr, which also show no emission lines. Aperture biases in our spectroscopy are unlikely to explain the lack of observed star-formation in this group: the spectra sample the cen- tral ∼ 65h−2 kpc2 of these distant galaxies. Further sup- port for a lack of on-going star-formation in these systems is shown by the uniform red colors of those galaxies for which we have imaging in two passbands with WFPC2. We quantify the occurrence of passive late-type galaxies, and compare cluster and field populations, in Fig. 11. Us- ing the cumulative distribution of [Oii] 3727 EW, we find that for the morphological groups E and S0–Sb there is a significantly higher fraction of galaxies showing little or no [Oii] emission in clusters as compared to the field. The likelihood, P , that the cluster and field samples are drawn from the same population is less than log10 P < −2.4 for both E and S0–Sb samples. However, the comparison of the [Oii] distribution of the latest-type systems (Sc–Irr, T=7–10) shows no significant difference between the clus- ter and field, although the number of galaxies is somewhat smaller. As an overall trend, then, there seems to be a decline in current star-formation at a fixed Hubble-type from field to cluster (see also Balogh et al. 1998). Furthermore, based on [Oii] EW alone as a measure of star formation, we see no evidence for enhanced star-formation in gas-rich cluster galaxies compared to the equivalent morphological sample in the field. We discuss this incidence of passive late-type galaxies further in P98. Fig. 12. A comparison of the distribution of D4000 measures in the different morphological types, for both cluster and field galaxies. In contrast to these results based on [Oii] EW, the dis- tribution of D4000 strengths (Fig. 12) is very similar for cluster and field: the individual morphological types are indistinguishable in D4000 at better than log10 P > −1 in each case. Thus, while [Oii], the tracer of current star- formation, shows a decline in the cluster, this does not appear to be reflected in an index sensitive to the star- formation averaged over a somewhat longer period of the recent past (∼ 1–3 Gyrs). 5. RESULTS AND DISCUSSION 5.1. The Incidence of k+a/a+k and e(a) Galaxies Our spectral catalog exhibits one effect that is especially strong: the incidence of k+a/a+k galaxies in distant clus- ters is very high compared to the surrounding field. Ta- ble 7 shows that in the cluster sample we have 60 examples of k+a and 18 examples of a+k, totaling 18% of the sam- ple. This is similar to the typical value of ∼10–20% found by magnitude-limited surveys of distant clusters (DG92; 14Caldwell & Rose (1997) have reported a frequency of ∼15% of notably stronger Balmer lines in early type galaxies in five low-redshift clusters. These are for the most part lower luminosity systems, with Hδ < 3.0 A, which the authors suggest are the remnants of earlier bursts. The results of that study do not, therefore, conflict with the much lower frequency found by Dressler and Shectman for stronger, more luminous systems. Dressler et al. 13 Couch et al. 1998). However, this value strongly contrasts with the 7 occurrences, all k+a, found in the high-redshift field sample, only 2%. Indeed, 4 of these 7 cases are either uncertain or border- line, a far greater fraction than for the cluster sample, so an incidence of ∼1% is compatible with these data. For the low-redshift Las Campanas Redshift Survey (hereafter LCRS), Zabludoff et al. (1996) found an incidence of 0.2%, but their selection criteria included a stronger limit on Hδ of 5.5 A and they note that the number increases to 0.6% when the limit is dropped to 4.5 A. Hashimoto (1998) has evaluated the occurrence of the spectral classes as defined in this paper for the LCRS, and finds 2.3% for the occur- rence of k+a/a+k types. In summary, these data seem to point to at most a factor two increase in the frequency of k+a/a+k types between the low- and intermediate- redshift field populations. This is in marked contrast to the order-of-magnitude increase in the frequency of k+a types in rich clusters. At low redshift, this frequency is ∼ 1% (determined using the Dressler and Shectman (1988) < catalog), compared to the 18% found here for the z ∼ 0.5 clusters.14 Zabludoff et al. attributed many of the low-redshift field “E+A’s” as due to mergers and strong interactions, since morphologies of this type are often observed in the low- redshift examples. The expected evolution can be esti- mated from the change in the incidence of close pairs (Zepf & Koo 1989; Patton et al. 1997), which would be predicted to be the parent population. Patton et al. (1997) claim that the proportion of close pairs (two galaxies within 20h−1 kpc) increases by a factor of ∼ 1.5 between z = 0 and z = 0.33. Extrapolating this behavior to < z >= 0.42 would predict an increase in the fraction of close pairs of ∼ 2–3 over that seen locally. Although we see at most a factor of two increase in the k+a population from low to high redshift using the LCRS and our sample, this does not rule out that a significant fraction of field k+a’s are due to such mergers. Zabludoff et the as that, further argue al. merger/interaction mechanism appears to be responsi- ble for low-redshift field examples of such galaxies, it is reasonable to conclude that mergers may also be responsi- ble for the k+a/a+k galaxies in the intermediate-redshift clusters. However, the radically different evolution de- scribed above of the k+a/a+k population between cluster and field environments suggests that the cluster environ- ment is crucial in either the formation of cluster k+a/a+k galaxies, or in prolonging their visibility. This could in part be due to an increased propensity for mergers in the groups infalling into the intermediate-redshift clusters. However, our morphological analysis (S97) finds only a minority of cases of k+a spectra where the galaxy shows signs of a classic two-body merger, as Zabludoff et al. found for the low-redshift field examples. We conclude, then, that at least one mechanism other than mergers is responsible for the large fraction of k+a/a+k galaxies in intermediate-redshift clusters. As we discuss in B98, the ma jority of k+a/a+k spec- tra are the result of a sudden decline in star formation 14 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies rate that followed a substantial rise, or burst, of star- formation, leaving a population of A-stars to dominate the light for ∼ 109 years. Given the generic nature of the star-formation history required to form an a+k/k+a, mergers are obviously not a unique explanation for the k+a/a+k phenomena. For example, accretion of smaller satellites, instead of mergers of comparable mass systems, is not inconsistent with the morphologies we see. The greater fraction of a+k/a+k galaxies in the intermediate- redshift clusters as compared to the field is likely to be connected as well with the frequency of e(a) galaxies in these environments (B98). 5.2. The Distribution and Kinematic Properties of the Cluster Galaxies As a final exercise in the comparison of spectroscopic properties with other cluster characteristics, we examine the radial distributions of our cluster sample as a function of spectroscopic type. We begin by assigning field centers — these positions are given in Table 2. There is usually little ambiguity of this due to the presence of a D or cD galaxy, these have been confirmed as the cluster centers in all cases from our the weak lensing analysis in Smail et al. (1997b). Even in more complex cases, such as Cl 1447+26, the ambiguity in choosing a center will play little role over the large range in radius we investigate. 1 0.8 0.6 0.4 0.2 0 -2 -1.5 -1 -0.5 0 0.5 a much more extended distribution. What is perhaps more interesting is the way the k+a/a+k types, which may sen- sibly interpreted as post-starburst galaxies, avoid the cen- ters in contrast to the k types, but are far less extended than the emission-line galaxies.15 The near absence of k+a types in the field, discussed above, coupled with the sudden rise in their frequency as the cluster center is ap- proached, with an almost complete demise in the central regions, appears to be clear evidence for the environment effecting either their formation or visibility. We note in passing that a similar diagram subdividing the e types into e(a), e(c), and e(b) shows no significant difference, though there is a hint that the e(a) class has a slightly more extended distribution. We now investigate the rudimentary kinematics of the sample of cluster members. In Table 8 we list the restframe velocity dispersions and uncertainties for the entire clus- ter sample broken down in terms of morphological type, spectral class, disturbance and activity class (the latter three sections refer only to those galaxies lying within the WFPC2 fields). The distributions for the spectral and morphological types are also shown in Fig. 14. These val- ues are calculated using the mean cluster redshifts listed in Table 2 and are simple averages across the cluster (no allowance has been made for different velocity dispersion for the different clusters – when such corrections are ap- plied they make no qualitative change to the conclusions listed below). The uncertainties in the velocity dispersions are 1σ values estimated from bootstrap resampling of the observed distributions. Starting with the morphological samples in Table 8 we see a marked difference between the velocity dispersion of the elliptical galaxies and all the later-types, the latter having higher dispersions (including the S0 galaxies). A similar difference is noticeable when the sample is split into different spectral classes (now including the whole spec- troscopic catalog of members). Interestingly the galaxies whose spectra were too poor to be classified, the “?” class, show the lowest dispersion – suggesting that these may be predominantly passive, cluster galaxies. The strongest trend is the significantly higher velocity dispersion of the presently or recently star-forming systems compared to the passive population (c.f. Dressler 1986). In particular, combining the different spectral classes (e(all) comprises e(a)/e(b)/e(c)/e(n)/e) from Table 8 we find that the emission-line and k-type galaxies have relative disper- sions of σem /σk = 1.40 ± 0.16, with the k+a/a+k galaxies being intermediate between the two. The higher disper- sions of the active populations are consistent with these galaxies being less virialised than the k-type population. Such a trend can also be discerned in the variation of ve- locity with activity (as traced by the [Oii] EW) within the individual morphological types. Splitting each of the more active morphological classes (Sb-Sd/Irr) at its me- dian [Oii] EW into “low” and “high” activity samples we find the dispersions listed at the bottom of Table 8 for the different morphological samples. For all three morphologi- cal types, the more active sample shows the higher velocity dispersion. A higher velocity dispersion is often taken as a sign of an infalling population, but, as we discuss below, Fig. 13. The cumulative radial distribution of different spectral types. These are shown for the all members from the whole sample which have MV < −19 + 5 log10 h. There is a clear difference between the radial distribution of k, k+a/a+k, and e type galaxies, with the former being most concentrated, the latter the least. The k+a/a+k class seems to be intermediate between the two showing a similar decline to the k types on the outskirts of the cluster, but a flatter distribution in the core, more in keeping with that seen for the e types. In Fig. 13 we show for the combined clusters the cumu- lative radial distribution for different spectroscopic types. This procedure is crude because it averages over the non- spherical distribution of galaxies within the clusters, but it may provide some insight into the characteristic distribu- tions of different classes of galaxies. Not surprisingly, the k types, generally made up of E and S0 galaxies (Fig. 10), but including significant early-type spirals as well, are the most concentrated population in these clusters (c.f. S97). Also, not surprisingly, the emission line galaxies strongly avoid the center (r ≤ 50h−1 kpc) of these clusters and have 15 It is tempting to describe this distribution as a “thick shell”, but we consider this potentially misleading due to the substantial departures from spherical symmetry exhibited by our clusters. Rather, it is probably more instructive to think of k+a/a+k types occurring most frequently at an intermediate radius R ∼ 200kpc. Dressler et al. 15 Velocity Dispersions of Cluster Populations Table 8 Sample N E S0 Sa Sb Sc Sd/Irr k k+a a+k e(a) e(c) e(b) e(n) e ? k k+a/a+k e(all) D = 0 D = 1 D = 2 D = 3 D = 4 Sb-high Sb-low Sc-high Sc-low Sd-high Sd-low 50 24 39 38 29 20 186 60 18 44 64 20 7 13 12 186 78 148 100 54 34 12 4 10 28 14 15 10 10 σ < v > (km s−1 ) (km s−1 ) Morphological Types 974.9 −109.9 1709.7 −378.7 1336.5 34.3 1290.6 179.5 640.9 1212.6 1894.6 −323.2 Spectral Classes 12.5 105.4 120.9 −27.2 −226.4 −168.8 576.1 743.2 367.6 12.5 109.0 −36.3 1064.8 1421.6 1236.4 1420.9 1437.1 1740.1 1495.3 1400.1 694.1 1064.8 1373.4 1486.0 Disturbance Classes 1252.9 −33.9 1382.8 207.6 58.9 1417.5 1572.1 307.5 −681.9 2884.3 Activity Classes ... ... ... ... ... ... 1429.5 1265.5 1293.7 1084.2 2274.5 1243.1 δσ (km s−1 ) 78.7 374.1 137.2 147.7 219.2 231.9 55.2 204.9 223.7 151.3 126.2 260.9 404.9 288.8 116.3 55.2 172.4 90.1 144.9 146.2 202.3 329.4 807.8 294.1 173.6 334.2 246.1 332.7 238.1 Fig. 14. a) The restframe velocities of cluster members, calculated relative to their respective cluster means (Table 2). These histograms are shown for the different spectroscopic classes. No scaling has been applied when combining the distributions from different clusters. b) The restframe velocities of cluster members, relative to their respective cluster means. These histograms are shown for the different morphological types. 16 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies including spatial information in our analysis shows only weak evidence for infall. We note that the [Oii] EW distributions for these ac- tive cluster members do not show any enhanced activ- ity over that seen in the surrounding field for any given morphological type (c.f. §4.2). Apparently, then, the ob- served correlation of velocity dispersion and activity (as measured by [Oii] EW) is not triggered by a mechanism which causes the higher star-formation rates due to the high relative velocities of the galaxies within the clusters (i.e. ram-pressure induced star-formation). We suggest in- stead that the correlation between activity and velocity dispersion reflects a decline in star-formation in the galax- ies which runs in parallel with and is causally linked their virialization within the clusters. In this regard we also mention the trend for more disturbed galaxies to have higher velocity dispersions, Table 8, a result which remains when we restrict the analysis to late-type galaxies, Sb–Irr. We next combine the velocity and positional information for the entire sample of clusters, focusing on the spectral classes, to calculate the velocity dispersion as a function of spectral class and radius. Dressler (1986) used the Gio- vanelli and Haynes (19ZZ) catalog of spirals in nearby clus- ters to show that gas-poor spirals tend to travel on radial orbits that take them into the cluster center, as compared to the more isotropic orbits of the gas-rich spirals. Here we divide the k, k+a/a+k, and emission-line classes into three bins of radial position, each containing one-third of their respective samples. The resulting velocity dispersions are shown in Fig. 15. 2000 1500 1000 500 0 0 0.2 0.4 0.6 0.8 Fig. 15. The velocity dispersions of the different spectral types, aver- aged over the entire sample, as a function of radius. The velocity disper- sion is everywhere higher for active systems compared to passive galaxies. Note that the k+a/a+k types exhibit a peak in velocity dispersion that may be related to their distinctive spatial distribution. Unlike the clear difference in the orbital properties of gas-deficient systems in nearby clusters, our sample ex- hibits ambiguous evidence at best. The k and e types both have velocity dispersion that falls gently with radius or are, within the errors, flat. This suggests populations on mildly radial orbits, possibly an infalling population, or a simple isothermal distribution. More puzzling is the peak in velocity dispersion for the k+a/a+k types, which does appear to be statistically significant. It is possible, of course, that higher velocities increase the chance of pro- ducing a k+a/a+k. It is also possible that this kinematic feature is connected with their unusual radial distribution, as mentioned above. A system of largely circular orbits that might characterize this distribution, which is concen- trated like the k-types but avoids the core, would appear to have a higher velocity dispersion due to pro jection of what are largely tangential velocities. This is, however, not consistent with the idea that such galaxies derive from an infalling population on what are basically radial orbits. At this point, the statistics are poor enough, and the range of models so broad, that it is not worthwhile to explore this further here. 6. CONCLUSIONS • We have presented detailed spectroscopic observations of 657 galaxies in the fields of 10 z = 0.37–0.56 clusters. Combining these with our detailed HST-based morpho- logical catalogs in these fields we construct samples of 204 cluster members and 71 field galaxies with both accurate spectral and morphological information. • Using observational and theoretical justifications we have constructed a new quantitative spectral classification scheme and use this to interpret correlations between our spectral information and other properties of the galaxies in our catalog. • Based upon an analysis of the [Oii] EW distributions, we find no evidence for an increase in the occurrence of strongly star-forming galaxies in the moderate-redshift cluster environment compared to the moderate-redshift field using morphologically-selected samples. However, we do find a large population of late-type cluster, but not field, galaxies which show little or no evidence of on-going star-formation. • This passive, late-type cluster population is related to our spectral classes k+a/a+k, both of which we interpret as indicative of post-starburst behavior. Galaxies with k+a/a+k spectra are an order-of-magnitude more frequent in the cluster environment compared to the high redshift field. • These k+a/a+k galaxies avoid the central regions of the clusters, in contrast to the k types, but are also far less extended than the emission-line galaxies, and much less common in the field. This appears to be clear evidence for the environment effecting either their formation or visibil- ity. • A detailed analysis of the spectroscopic and morpholog- ical information discussed here will be presented in Pog- gianti et al. (1998). ACKNOWLEDGEMENTS We thank Ray Lucas at STScI for his enthusiastic help which enabled the efficient gathering of these HST observa- tions. BMP and HB warmly thank Steve Maddox for cru- cial help during the 1995 WHT run and in the subsequent reduction of those data. We also thank Alfonso Arag´on- Salamanca, Nobuo Arimoto and Amy Barger for useful dis- cussions and assistance. AD and AO acknowledge support from NASA through STScI grant 3857. IRS acknowledges support from a PPARC Advanced Fellowship and from Royal Society and Australian Research Grants while an Honorary Visiting Fellow at UNSW. WJC acknowledges support from the Australian Department of Industry, Sci- ence and Technology, the Australian Research Council and Sun Microsystems. This work was supported in part by the Formation and Evolution of Galaxies network set up by the European Commission under contract ERB FMRX- CT96-086 of its TMR program. We acknowledge the avail- ability of the Kennicutt’s (1992) atlas of galaxies from the NDSS-DCA Astronomical Data Center. Dressler et al. 17 REFERENCES Allington-Smith, J. R., et al., 1994, PASP, 106, 983 Abraham, R., et al., 1996, ApJ, 471, 694 Arag´on-Salamanca, A., Ellis, R. S., Couch, W. J., & Carter, D., 1993, MNRAS, 262, 764 Balogh, M. L., Schade, D., Morris, S. L., Yee, H. K. C., Carlberg, R. G., & Ellingson, E., 1998, ApJL, 504, L75. Barger, A. J., Arag`on-Salamanca, A., Ellis, R. S., Couch, W. J., Smail, I., & Sharples, R. M., 1996, MNRAS, 279, 1 Barger, A. J., Arag`on-Salamanca, A., Smail, I., Ellis, R. S., Couch, W. J., Dressler, A., Oemler, A. Jr, Butcher, H., & Sharples, R. M., 1998, ApJ, in press. Bothun, G. D., & Gregg, M. D., 1990, ApJ, 350, 73 Butcher, H., & Oemler, A. Jr, 1978, ApJ, 279, 18 Butcher, H., & Oemler, A. Jr, 1984, ApJ, 285, 426 Caldwell, N., & Rose, J. A., 1997, AJ, 113, 492 Couch, W. J., & Sharples, R. M., 1987, MNRAS, 229, 423 (CS87) Couch, W. J., Ellis, R. S., Sharples, R. M., & Smail, I., 1994, ApJ, 430, 121 Couch, W. J., Barger, A. J., Smail, I., Ellis, R. S., Sharples, R. M., & Smail, I., 1998, ApJ, 497, 188 Dressler, A., 1986, in Nearly Normal Galaxies, ed. S. M. Faber, pp. 276. Dressler, A., & Shectman, S.A., 1987, AJ, 94, 899. Dressler, A., & Shectman, S.A., 1988, AJ, 95, 284. Dressler, A., & Gunn, J. E., 1983, ApJ, 270, 7. Dressler, A., & Gunn, J. E., 1992, ApJS, 70, 1 (DG92) Dressler, A., Oemler, A. Jr., Butcher, H., & Gunn, J. E., 1994, ApJ, 430, 107 Dressler, A., Oemler, A. Jr., Couch, W. J., Smail, I., Ellis, R. S., Barger, A., Butcher, H., Poggianti, B. M., & Sharples, R. M., 1997, ApJ, 490, 577. Ellis, R. S., Smail, I., Dressler, A., Couch, W. J., Oemler, A., Butcher, H., & Sharples, R. M., 1997, ApJ, 483, 582 Franx, M., 1993, ApJL, 407, L5 Fisher, D., Fabricant, D., Franx, M., van Dokkum, P., 1998, ApJ, in press Giovanelli, & Haynes, 19ZZ, ZZZ Hashimoto, Y., Oemler, A. Jr, 1999, ZZZ Hashimoto, Y., 1998, private communication. Kennicutt, R. C. Jr, 1992, ApJS, 79, 255 Kells, W., Dressler, A., Sivaramakrishna, A., Carr, D., Koch, E., Epps, H., et al., 1998, PASP, submitted. Lavery, R. J., & Henry, J. P., 1994, ApJ, 426, 524 Lubin, L. M., 1996, AJ, 112, 23 Oemler, A., Dressler, A., & Butcher, H. R., 1997, ApJ, 474, 561 Patton, D.R., Pritchet, C.J., Yee, H.K.C., Ellingson, E., Carlberg, R.G., 1997, ApJ, 475, 29 Poggianti, B. M. et al., 1998, in preparation (P98) Rakos, K. D., & Schombert, J. M., 1995, ApJ 439, 47 Rakos, K. D., Odell, A. P., & Schombert, J. M., 1997, ApJ, 490, 194 Smail, I., Ellis, R. S., Dressler, A., Couch, W. J., Oemler, A. Jr, Butcher, H., & Sharples, R. M., 1997a, ApJ, 479, 70 Smail, I., Dressler, A., Couch, W. J., Ellis, R. S., Oemler, A. Jr, Butcher, H., & Sharples, R. M., 1997b, ApJS, 110, 213 (S97) Smail, I., Edge, A. C., Ellis, R. S., & Blandford, R. D., 1998, MN- RAS, 293, 124 Stanford, S. A., Eisenhardt, P. R., & Dickinson, M. E., 1995, ApJ, 450, 512 Stanford, S. A., Eisenhardt, P. R., & Dickinson, M. E., 1998, ApJ, in press. Zabludoff, A.I., Zaritsky, D., Lin, H., Tucker, D., Hashimoto, Y., Shectman, S.A., Oemler, A. Jr., Kirsher, R.P., 1996, ApJ, 466, 104. Zepf, S.E., Koo, D.C., 1989, ApJ, 337, 34 Zijlstra, A., Giraud, E., Melnick, J., Dekker, H., & D’Odorico, S., 1996, ESO Operating Manual, version 3.0, ESO. Table 4A16 Spectroscopic Catalog of Cluster Galaxies Cluster ID (2) z (3) q (4) [Oii] (5) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 0.3910 0.3990 0.3825 0.3900 0.3910 0.3900 0.3755 0.3830 0.3860 0.3950 0.3970 0.3960 0.3900 0.3900 0.3956 0.3990 0.3902 0.3830 0.3940 0.3850 . . . 0.0 4 −50.2 2 0.0 3 0.0 4 −14.4 2 0.0 2 0.0 1 0.0 3 ... 3 ... 1 −43.3 3 0.0 3 0.0 3 0.0 3 0.0 2 −20.7 2 −4.9 1 2 0.0 1 −137.3 0.0 1 . . . Hδ (6) ... ... 2.3 ... ... ... ... ... ... 3.6: 3.0 ... ... ... ... ... 3.6 ... ... ... D4000 (7) Class (8) δRA (9) δDec (10) IDHST (11) X (12) Y (13) Morph (14) T (15) D (16) 2.52 1.21 2.13 2.05 1.28 1.56 1.80 1.88 1.85 2.06 1.25 2.01 2.10 2.08 2.02 1.49 1.64 2.08 0.81 2.08 . . . 69.1 −64.4 k: 97.4 11.8 e(b),CSB 60.1 −15.3 k 27.6 −15.2 k: 7.2 42.3 e(c)/e(a) 31.7 −5.5 k 0.9 −44.0 k 15.5 −14.4 k: 12.0 −16.9 e(c): k+a/k 46.7 38.6 e(b),CSB −26.7 −46.9 −11.9 −24.2 k: 2.8 4.9 k k: 0.0 0.0 −30.1 −22.7 k 23.9 e(c): 53.6 −30.9 −18.9 e(c) k: −15.6 33.5 e(n),CSB −70.2 −40.2 −18.9 k: 39.3 . . . 294 871 739 573 659 577 403 461 460 ... 203 304 ... 343 133 ... 127 147 ... 113 . . . 1317 1318 1198 1122 1373 1224 777 1100 1069 ... 689 942 ... 1204 914 ... 949 1499 ... 1552 E Scd Sc E Sd S0 S0/E S0/a Sb ... Sd E ... E S0 ... Sbc E ... E 343 1365 1008 692 785 712 497 572 543 ... 234 327 ... 387 146 ... 128 158 ... 114 . . . 0 1 2 1 1 0 0 0 3 ... 2 0 ... 0 0 ... 0 1 ... 0 −5 6 5 −5 7 −2 −3 0 3 ... 7 −5 ... −5 −2 ... 4 −5 ... −5 . . . Int (17) T? I? I M? ... ... ... I I/M ... T ... ... ... ... ... ... I? ... I? . . . Maga,b (18) Cola,c Magd DG (19) (20) Cold DG (21) 19.96 20.37 20.12 18.35 19.97 19.51 19.35 20.08 19.84 ... 20.32 18.47 ... 18.35 19.94 ... 19.45 19.41 ... 18.89 3.30 1.60 2.56 3.42 1.95 3.18 3.32 2.88 3.32 ... 1.49 3.42 ... 3.51 3.32 ... 2.77 3.29 ... 3.40 . . . 21.54 21.44 21.01 19.77 21.30 20.86 20.66 21.49 21.45 20.26 21.48 19.63 19.29 19.49 21.10 21.34 20.73 20.79 21.87 19.73 1.66 0.39 1.10 1.58 0.76 1.26 1.45 1.38 1.47 1.48 0.37 1.58 1.61 1.71 1.42 0.84 0.96 1.64 0.36 1.58 . . . Run (22) DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 DG92 Mask (23) Features (24) Commen (25) ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... DG16 ... DG19; ... DG59 ... DG101 ... DG104; ... DG106 ... DG112 ... DG120 ... DG123; ... DG128;Hd H,K,G,Mg DG138; ... DG140 ... DG144 ... DG148 ... DG175 ... DG177; ... DG178 ... ... DG196 OII,Hb,OIII D G198; ... DG210 Table 4B17 Spectroscopic Catalog of Field Galaxies Cluster ID (2) 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 z (3) q (4) [Oii] (5) 0.1393 0.4545 0.2132 0.2123 0.4758 0.6547: 0.2252 0.2279 0.5558 0.7128 0.8089 0.6569 0.6564 0.6946 0.1110 0.2153 0.2478 0.1111 0.2129 0.2724 . . . 3 ... 3 −12.9 2 0.0 3 −19.3 2 −43.8 4 0.0 3 −22.8 2 −51.8 2 −27.2 2 0.0 4 −15.9 −3.6 2 2 0.0 2 −37.8 2 −35.7 1 0.0 2 −53.4 2 −46.0 1 −29.5 0.0 1 . . . Hδ (6) 3.1 0.0 7.4 4.8 6.5 0.0 4.6 0.0 4.7 0.0 0.0 0.0 0.0 3.9 4.7 0.0 2.3 5.0 1.3 0.0 D4000 (7) Class (8) δRA (9) δDec (10) IDHST (11) X (12) Y (13) Morph (14) T (15) D (16) Int (17) Maga,b (18) Cola,c Magd DG (19) (20) Cold DG (21) ... 1.41 2.11 1.74 1.54 1.80 1.42 1.46 1.35 1.77 1.55 1.59 1.92 1.22 2.17 1.85 1.63 1.49 1.81 1.93 . . . −65.2 28.9 e(c) 43.6 −28.4 e(c),CSB 57.6 −76.7 e(a) −6.3 −35.4 e(a) 66.7 168.3 e(a) 43.6 228.3 k −44.8 e(a) 281.9 31.3 e(b),CSB −137.0 17.1 −1.0 e(a) 15.0 133.4 k+a: 45.7 177.0 e(c) 29.9 200.6 e(c) 43.6 228.4 k 32.9 e(a) 246.1 e(a),CSB −121.6 29.9 −64.7 −254.1 k −58.7 −38.6 e(b) 29.8 −121.6 e(b),e(a) 38.9 −57.4 e(c) −53.9 e(c): 86.8 . . . 797 ... ... ... ... ... ... ... 302 ... ... ... ... ... ... ... ... ... ... ... . . . 1221 ... ... ... ... ... ... ... 1371 ... ... ... ... ... ... ... ... ... ... ... Scd ... ... ... ... ... ... ... Scd ... ... ... ... ... ... ... ... ... ... ... 1103 ... ... ... ... ... ... ... 341 ... ... ... ... ... ... ... ... ... ... ... . . . 0 ... ... ... ... ... ... ... 2 ... ... ... ... ... ... ... ... ... ... ... 6 ... ... ... ... ... ... ... −5 ... ... ... ... ... ... ... ... ... ... ... . . . ... ... ... ... ... ... ... ... I? ... ... ... ... ... ... ... ... ... ... ... . . . 17.90 ... ... ... ... ... ... ... 20.18 ... ... ... ... ... ... ... ... ... ... ... 2.57 ... ... ... ... ... ... ... 2.82 ... ... ... ... ... ... ... ... ... ... ... . . . 19.20 21.17 21.30 21.97 22.51 ... ... 23.31 ... 22.20 ... ... ... ... 21.44 ... ... ... 19.31 20.40 0.73 0.63 0.79 0.84 0.67 ... ... 0.17 ... 1.52 ... ... ... ... 0.29 ... ... ... 0.59 1.27 . . . Run (22) DG92 DG92 P9408 P9408 P9408 P9408 P9408 P9408 P9412 P9412 P9412 P9412 P9412 P9412 P9412 P9410 P9410 P9410 P9410 P9410 Mask (23) Features (24) Commen (25) Ha,SII ... OII,OIII ... ew1 11 H,K,Bal,Ha ew1 13 OII,Hb,abs ew1 27 OII,Bal ew1 32 H,K,G,Mg ew1 37 OII,OIII,Ha ew1 7 OII,Hb,OIII,Ha ew2 15 OII,OIII,Bal ew2 26 H,K,G ew2 28 OII,H,K ew2 29 H,K,Mg,OII ew2 30 H,K,G,Mg ew2 32 OII,H,K,Mg,Bal ew2 5 OII,OIII,Ha ns1 1 H,K,G,Hb,Mg OII,Hb,OIII,Ha ns1 11 ns1 19 OII,Hb,OIII,Ha ns1 20 OII,Hb,OIII,Ha ns1 24 H,K,G DG47; DG220; DG290; DG185; DG13; noisy Ha= DG325; DG10 p w strong DG312 Hd str Ha= DG257; DG289; 20 A Spectroscopic Catalog of 10 Distant Rich Clusters of Galaxies Fig. 8a. The images of those galaxies in our sample lying within the WFPC2 frame, grouped into spectroscopic classes. This first panel shows those galaxies with k spectral types. Each image is 5′′ × 5′′ (or 15.5–18.6 h−1 kpc depending upon the cluster’s redshift) and has the same orientation as the HST field (S97). The cluster and galaxy ID from the WFPC2 catalogs (from Tables 4 in S97, or Table 4a below) and the spectral class and morphological type are marked on each frame. For the plates in Fig. 8 see http://www.ociw.edu/∼irs/morphs2.html#figs Fig. 8a. continued. Fig. 8a. continued. Fig. 8b. The k+a sample. Fig. 8c. The a+k sample. Fig. 8d. The e(a) sample. Fig. 8e. The e(c) sample. Fig. 8f. The e(b) sample. Fig. 8g. The e(n) sample.
0801.1403
1
0801
2008-01-09T11:19:45
The importance of magnetic-field-oriented thermal conduction in the interaction of SNR shocks with interstellar clouds
[ "astro-ph" ]
We explore the importance of magnetic-field-oriented thermal conduction in the interaction of supernova remnant (SNR) shocks with radiative gas clouds and in determining the mass and energy exchange between the clouds and the hot surrounding medium. We perform 2.5D MHD simulations of a shock impacting on an isolated gas cloud, including anisotropic thermal conduction and radiative cooling; we consider the representative case of a Mach 50 shock impacting on a cloud ten-fold denser than the ambient medium. We consider different configurations of the ambient magnetic field and compare MHD models with or without the thermal conduction. The efficiency of the thermal conduction in the presence of magnetic field is, in general, reduced with respect to the unmagnetized case. The reduction factor strongly depends on the initial magnetic field orientation, and it is minimum when the magnetic field is initially aligned with the direction of shock propagation. The thermal conduction contributes to suppress hydrodynamic instabilities, reducing the mass mixing of the cloud and preserving the cloud from complete fragmentation. Depending on the magnetic field orientation, the heat conduction may determine a significant energy exchange between the cloud and the hot surrounding medium which, while remaining always at levels less than those in the unmagnetized case, leads to a progressive heating and evaporation of the cloud. This additional heating may contrast the radiative cooling of some parts of the cloud, preventing the onset of thermal instabilities.
astro-ph
astro-ph
Draft version October 30, 2018 Preprint typeset using LATEX style emulateapj v. 08/13/06 THE IMPORTANCE OF MAGNETIC-FIELD-ORIENTED THERMAL CONDUCTION IN THE INTERACTION OF SNR SHOCKS WITH INTERSTELLAR CLOUDS S. Orlando1, F. Bocchino1 INAF - Osservatorio Astronomico di Palermo "G.S. Vaiana", Piazza del Parlamento 1, 90134 Palermo, Italy Dip. di Scienze Fisiche & Astronomiche, Univ. di Palermo, Piazza del Parlamento 1, 90134 Palermo, Italy F. Reale2,1, G. Peres2,1 and P. Pagano Draft version October 30, 2018 ABSTRACT We explore the importance of magnetic-field-oriented thermal conduction in the interaction of su- pernova remnant (SNR) shocks with radiative gas clouds and in determining the mass and energy exchange between the clouds and the hot surrounding medium. We perform 2.5D MHD simulations of a shock impacting on an isolated gas cloud, including anisotropic thermal conduction and radiative cooling; we consider the representative case of a Mach 50 shock impacting on a cloud ten-fold denser than the ambient medium. We consider different configurations of the ambient magnetic field and compare MHD models with or without the thermal conduction. The efficiency of the thermal conduc- tion in the presence of magnetic field is, in general, reduced with respect to the unmagnetized case. The reduction factor strongly depends on the initial magnetic field orientation, and it is minimum when the magnetic field is initially aligned with the direction of shock propagation. The thermal conduction contributes to suppress hydrodynamic instabilities, reducing the mass mixing of the cloud and preserving the cloud from complete fragmentation. Depending on the magnetic field orientation, the heat conduction may determine a significant energy exchange between the cloud and the hot sur- rounding medium which, while remaining always at levels less than those in the unmagnetized case, leads to a progressive heating and evaporation of the cloud. This additional heating may contrast the radiative cooling of some parts of the cloud, preventing the onset of thermal instabilities. Subject headings: conduction -- magnetohydrodynamics -- shock waves -- ISM: clouds -- ISM: magnetic fields -- ISM: supernova remnants 1. INTRODUCTION The interaction of the shock waves of supernova rem- nants (SNRs) with the magnetized and inhomogeneous interstellar medium (ISM) is responsible of the great morphological complexity of SNRs and certainly plays a major role in determining the exchange of mass, momen- tum, and energy between diffuse hot plasma and dense clouds or clumps. These exchanges may occur through, for example, hydrodynamic ablation and thermal con- duction and, among other things, lead to the cloud crush- ing and to the reduction of the Jeans mass causing star formation. The propagation of hot SNR shock fronts in the ISM and their interaction with local over-dense gas clouds have been investigated with detailed hydrodynamic and MHD modeling. The most complete review of this problem in the unmagnetized, non-conducting, and non- radiative limits is provided by Klein et al. (1994). These studies have shown that the cloud is disrupted by the ac- tion of both Kelvin-Helmholtz (KH) and Rayleigh-Taylor (RT) instabilities after several crushing times, with the cloud material expanding and diffusing into the ambient medium. An ambient magnetic field can both act as a confinement mechanism of the plasma and be modified by the interstellar flow and by local field stretching. Also, a strong magnetic field is known to limit hydrodynamic instabilities developing during the shock-cloud interac- 1 Consorzio COMETA, via Santa Sofia 64, 95123 Catania, Italy 2 INAF, Viale del Parco Mellini 84, 00136 Roma, Italy tion by providing an additional tension at the interface between the cloud and the surrounding medium (e.g Mac Low et al. 1994; Jones et al. 1996). The interaction of the shock with a radiative cloud has been only recently analyzed in detail (e.g. Mellema et al. 2002; Fragile et al. 2004). 2D calculations have shown that the effect of the radiative cooling is to break up the clouds into numerous dense and cold fragments that survive for many dynamical timescales. In the case of the interaction between magnetized shocks and radiative clouds, the magnetic field may enhance the efficiency of the radiative cooling, influencing the size and distribu- tion of condensed cooled fragments (Fragile et al. 2005). The role played by the thermal conduction during the shock-cloud interaction has been less studied so far. In a previous paper, Orlando et al. (2005) (hereafter Paper I) have addressed this point in the unmagnetized limit. In particular, we have investigated the effect of thermal con- duction and radiative cooling on the cloud evolution and on the mass and energy exchange between the cloud and the surrounding medium; we have selected and explored two different physical regimes chosen so that either one of the processes is dominant. In the case dominated by the radiative losses, we have found that the shocked cloud fragments into cold, dense, and compact filaments sur- rounded by a hot corona which is ablated by the thermal conduction. Instead, in the case dominated by thermal conduction, the shocked cloud evaporates in a few dy- namical timescales. In both cases, we have found that 2 Orlando et al. the thermal conduction is very effective in suppressing the hydrodynamic instabilities that would develop at the cloud boundaries, preserving the cloud from complete de- struction. Orlando et al. (2006) and Miceli et al. (2006) have studied the observable effects of thermal conduction on the evolution of the shocked cloud in the X-ray band. Here, we extend the previous studies by investigat- ing the effect of the thermal conduction in a magne- tized medium, unexplored so far. Of special interest to us is to investigate the role of anisotropic thermal con- duction - funneled by locally organized magnetic fields - in the mass and energy exchange between ISM phases. In particular, we aim at addressing the following ques- tions: How and under which physical conditions does the magnetic-field-oriented thermal conduction influence the evolution of the shocked cloud? How do the mass mixing of the cloud material and the energy exchange between the cloud and the surrounding medium depend on the orientation and strength of the magnetic field and on the efficiency of the thermal conduction? To answer these questions, we take as representative the model case of a shock with Mach number M = 50 (corresponding to a post-shock temperature T ≈ 4.7 × 106 K for an unperturbed medium with T = 104 K) impacting on an isolated cloud ten-fold denser than the ambient medium. Paper I has shown that, in this case, the thermal conduction dominates the evolution of the shocked cloud in the absence of magnetic field. Around this basic configuration, we perform a set of MHD simu- lations, with different interstellar magnetic field orienta- tions, and compare models calculated with thermal con- duction turned either "on" or "off" in order to identify its effects on the cloud evolution. The paper is organized as follows: in Sect. 2 we de- scribe the MHD model and the numerical setup; in Sect. 3 we discuss the results; and finally in Sect. 4 we draw our conclusions. 2. THE MODEL We model the impact of a planar supernova shock front onto an isolated gas cloud. The shock propagates through a magnetized ambient medium and the cloud is assumed to be small compared to the curvature radius of the shock3. The fluid is assumed to be fully ionized with a ratio of specific heats γ = 5/3. The model in- cludes radiative cooling, thermal conduction (including the effects of heat flux saturation) and resistivity effects. The shock-cloud interaction is modeled by solving nu- merically the time-dependent non-ideal MHD equations (written in non-dimensional conservative form): ∂ρ ∂t + ∇ · (ρu) = 0 , ∂ρu ∂t + ∇ · (ρuu − BB) + ∇P∗ = 0 , (1) (2) ∂ρE ∂t + ∇ · [u(ρE + P∗) − B(u · B)] = ∇ · [B × (η∇ × B)] − ∇ · Fc − nenHΛ(T ) ,(3) 3 In the case of a small cloud, the SNR does not evolve signif- icantly during the shock-cloud interaction, and the assumption of a planar shock is justified (see also Klein et al. 1994). ∂B ∂t where + ∇ · (uB − Bu) = −∇ × (η∇ × B) , (4) P∗ = P + B2 2 , E = ǫ + 1 2 u2 + 1 2 B2 ρ , are the total pressure, and the total gas energy (internal energy, ǫ, kinetic energy, and magnetic energy) respec- tively, t is the time, ρ = µmH nH is the mass density, µ = 1.26 is the mean atomic mass (assuming cosmic abundances), mH is the mass of the hydrogen atom, nH is the hydrogen number density, u is the gas velocity, T is the temperature, B is the magnetic field, η is the resis- tivity according to Spitzer (1962), Fc is the conductive flux, and Λ(T ) represents the radiative losses per unit emission measure (e.g. Raymond & Smith 1977; Mewe et al. 1985; Kaastra & Mewe 2000). We use the ideal gas law, P = (γ − 1)ρǫ. In order to track the original cloud material, we use a tracer that is passively advected in the same manner as the density. We define Ccl the mass fraction of the cloud inside the computational cell. The cloud material is initialized with Ccl = 1, while Ccl = 0 in the ambient medium4. During the shock-cloud evolution, the cloud and the ambient medium mix together, leading to regions with 0 < Ccl < 1. At any time t the density of cloud material in a fluid cell is given by ρcl = ρCcl. The thermal conductivity in an organized magnetic field is known to be highly anisotropic and it can be ex- traordinarily reduced in the direction transverse to the field. The thermal flux, therefore, is locally split into two components, along and across the magnetic field lines, Fc = Fk i + F⊥ j, where Fk = (cid:18) 1 [qspi]k + F⊥ = (cid:18) 1 [qspi]⊥ + , 1 [qsat]k(cid:19)−1 [qsat]⊥(cid:19)−1 1 (5) , to allow for a smooth transition between the classical and saturated conduction regime. In Eqs. 5, [qspi]k and [qspi]⊥ represent the classical conductive flux along and across the magnetic field lines (Spitzer 1962) [qspi]k = −κk[∇T ]k ≈ −5.6 × 10−7T 5/2 [∇T ]k [qspi]⊥ = −κ⊥[∇T ]⊥ ≈ −3.3 × 10−16 n2 H T 1/2B2 [∇T ]⊥ (6) where [∇T ]k and [∇T ]⊥ are the thermal gradients along and across the magnetic field, and κk and κ⊥ (in units of erg s−1 K−1 cm−1) are the thermal conduction coeffi- cients along and across the magnetic field lines5, respec- tively. The saturated flux along and across the magnetic field lines, [qsat]k and [qsat]⊥, are (Cowie & McKee 1977) 4 We checked that the used numerical scheme ensures that always 0 ≤ Ccl ≤ 1. 5 For the values of T , nH and B used here, κk/κ⊥ ≈ 1016 at the beginning of the shock-cloud interaction. Anisotropic thermal conduction in SNRs 3 TABLE 1 Summary of the initial physical parameters characterizing the MHD simulations. TABLE 2 Summary of the MHD simulations. In all runs the shock Mach number is M = 50, the density contrast is χ = 10, and the cloud crushing time is τcc ≈ 5.4 × 103 yr. Temperature Density Velocity ISM Cloud Post-shock medium: 104 K 103 K 4.7 × 106 K 0.1 cm−3 1.0 cm−3 0.4 cm−3 0.0 0.0 430 km s−1 [qsat]k = −sign(cid:0)[∇T ]k(cid:1) 5φρc3 [qsat]⊥ = −sign ([∇T ]⊥) 5φρc3 s , s , (7) where cs is the isothermal sound speed, and φ is a num- ber of the order of unity; we set φ = 0.3 according to the values suggested for a fully ionized cosmic gas: 0.24 < φ < 0.35 (Giuliani 1984; Borkowski et al. 1989; Fadeyev et al. 2002, and references therein). As dis- cussed in Paper I, this choice implies that no thermal precursor develops during the shock propagation, consis- tent with the fact that no precursor is observed in young and middle aged SNRs. The initial unperturbed ambient medium is magne- tized, isothermal (with temperature Tism = 104 K, cor- responding to an isothermal sound speed cism = 11.5 km s−1), and uniform (with hydrogen number density nism = 0.1 cm−3; see Table 1). The gas cloud is in pres- sure equilibrium with its surrounding and has a circular cross-section with radius rcl = 1 pc; its radial density distribution is given by ncl(r) = nism + ncl0 − nism cosh [σ (r/rcl)σ] , (8) where ncl0 is the hydrogen number density at the cloud center, r is the radial distance from the cloud center and σ = 10. The above distribution describes a thin tran- sition layer (∼ 0.3 rcl) around the cloud that smoothly brings the cloud density to the value of the surrounding medium6. The initial density contrast between the cloud center and the ambient medium is χ = ncl0/nism = 10. The cloud temperature is determined by the pressure bal- ance across the cloud boundary. The SNR shock front propagates with a velocity w = Mcism in the ambient medium, where M is the shock Mach number, and cism is the sound speed in the inter- stellar medium; we consider a shock propagating with M = 50, i.e. a shock velocity w ≈ 570 km s−1 and a temperature Tpsh ≈ 4.7 × 106 K. As discussed in Paper I, in this case (for a cloud with rcl = 1 pc and χ = 10) the cloud dynamics would be dominated by thermal con- duction in the absence of magnetic field. The post-shock conditions of the ambient medium well before the im- pact onto the cloud are given by the strong shock limit (Zel'dovich & Raizer 1966). Starting from this basic configuration, we consider a set of simulations with different initial magnetic field orien- tations. We adopt a 2.5D Cartesian coordinate system (x, y), implying that the simulated clouds are cylinders 6 A finite transition layer, in general, is expected in real interstel- lar clouds due, for instance, to thermal conduction (Balbus 1986; see also Nakamura et al. 2006). Run NN NR TN TR NN-Bx4 NN-By4 NN-Bz4 TN-Bx4 TN-By4 TN-Bz4 TR-Bx1 TR-By1 TR-Bz1 TR-Bx4 TR-By4 TR-Bz4 TR-Bx100 TR-By100 TR-Bz100 TR-Bz4-hr TR-Bz4-hr2 B µG 0 0 0 0 1.31 1.31 1.31 1.31 1.31 1.31 2.63 2.63 2.63 1.31 1.31 1.31 0.26 0.26 0.26 1.31 1.31 β0 ∞ ∞ ∞ ∞ 4 4 4 4 4 4 1 1 1 4 4 4 100 100 100 4 4 Field Comp. Therm. Cond. Rad. Losses Res.a − − − − Bx By Bz Bx By Bz Bx By Bz Bx By Bz Bx By Bz Bz Bz no no yes yes no no no yes yes yes yes yes yes yes yes yes yes yes yes yes yes no yes no yes no no no no no no yes yes yes yes yes yes yes yes yes yes yes 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 264 528 aInitial number of zones per cloud radius extending infinitely along the z axis perpendicular to the (x, y) plane. The primary shock propagates along the y axis. In this geometry, we consider three different field orientations: 1) parallel to the planar shock and per- pendicular to the cylindrical cloud, 2) perpendicular to both the shock front and the cloud, and 3) parallel to both the shock and the cloud. The magnetic field com- ponents along the x and the z axis are enhanced by a factor (γ + 1)/(γ − 1) (where γ is the ratio of specific heats) in the post-shock region (in the strong shock limit; Zel'dovich & Raizer 1966), whereas the component along the y axis is continuous across the shock. We include runs in the strong and weak magnetic field limits, consider- ing initial field strengths of B = 2.63, 1.31, 0.26, 0 µG in the unperturbed ambient medium7, corresponding to β0 = 1, 4, 100, ∞, where β0 = P/(B2/8π) is the ratio of thermal to magnetic pressure in the pre-shock region. This range of β0 includes typical values inferred for the diffuse regions of the ISM (e.g. Mac Low & Klessen 2004) and for shock-cloud interaction regions in evolved SNR shells (e.g Bocchino et al. 2000). There is no magnetic field component exclusively associated to the cloud. We follow the shock-cloud interaction for 3.5 τcc, where τcc ≈ χ1/2rcl/w is the cloud crushing time, i.e. the char- acteristic time of the shock transmission through the cloud; for the conditions considered here (χ = 10 and M = 50), τcc ≈ 5.4 × 103 yr. Each simulation is repeated either with or without thermal conduction for each field orientation. Table 2 lists the runs and the initial physical parameters of the simulations. We solve numerically the set of MHD equations us- 7 The unmagnetized case (i.e. B = 0) described here is analo- gous to the one studied in Paper I except for the fact that in the present case the cloud is a cylinder rather than a sphere and has smooth boundaries. 4 Orlando et al. Fig. 1. -- Mass density distribution (gm cm−3) in the (x, y) plane, in log scale, in the simulations NN (left half panels) and TR (right half panels), sampled at the labeled times in units of τcc. The contour encloses the cloud material. ing flash (Fryxell et al. 2000), a multiphysics code in- cluding the paramesh library (MacNeice et al. 2000) for the adaptive mesh refinement. The MHD equations are solved using the flash implementation of the HLLE scheme (Einfeldt 1988). The code has been extended with additional computational modules to handle the radiative losses and the anisotropic thermal conduction (see Pagano et al. 2007, for the details of the implemen- tation). The 2.5D Cartesian (x, y) grid extends between −4 and 4 pc in the x direction and between −1.4 and 6.6 pc in the y direction. Initially the cloud is located at (x, y) = (0, 0) and the primary shock front propagates in the direction of the y axis. At the coarsest resolution, the adaptive mesh algorithm used in the flash code uniformly cov- ers the 2.5D computational domain with a mesh of 42 blocks, each with 82 cells. We allow for 5 levels of refine- ment, with resolution increasing twice at each refinement level. The refinement criterion adopted (Lohner 1987) follows the changes of the density and of the tempera- ture. This grid configuration yields an effective resolu- tion of ≈ 7.6 × 10−3 pc at the finest level, corresponding to ≈ 132 cells per cloud radius. In Sect. 3.5, we discuss the effect of spatial resolution on our results, considering the additional runs TR-Bz4-hr and TR-Bz4-hr2 which use an identical setup to run TR-Bz4, but with higher resolution (≈ 264 and ≈ 528 cells per cloud radius, re- spectively; see Table 2). We use a constant inflow boundary condition for the post-shock gas at the lower boundary, with free outflow elsewhere. For runs with zero magnetic field (β0 = ∞), we use reflecting boundary conditions at x = 0 along the symmetry axis of the problem and only evolve half of the grid. 3. RESULTS 3.1. Dynamical evolution Figs. 1 and 2 show the evolution of the mass density in the (x, y) plane in the simulations with β0 = ∞ (runs NN and TR) and with β0 = 4 (runs NN-Bx4, NN-By4, NN- Bz4, TR-Bx4, TR-By4, TR-Bz4). The left (right) half panels show the result of models without (with) thermal conduction and radiative losses. From Fig. 1, we note that the thermal conduction drives the cloud evolution in the unmagnetized case (β0 = ∞; run TR): after the initial compression due to the primary shock, the cloud expands and gradually evaporates due to the heating driven by the thermal con- duction in a few dynamical timescales (see right half pan- els in Fig. 1). The heat conduction strongly contrasts the radiative cooling of some parts of the cloud and no ther- mal and hydrodynamic instabilities (visible in run NN; see left-panels in Fig. 1) develops during the cloud evo- lution, making the cloud more stable and longer-living (the mass mixing is strongly reduced; see Paper I for more details). We now discuss the effect of the magnetic-field-oriented (anisotropic) thermal conduction on the shock-cloud col- lision when an ambient magnetic field permeates the ISM. We first summarize the expected evolution in the presence of an ambient magnetic field, according to the well-established results of previous models without ther- mal conduction. We distinguish between fields perpen- dicular to the cylindrical clouds (i.e. with only Bx and By components; referred to as "external" fields by Frag- ile et al. 2005) and fields parallel to the cylindrical clouds (i.e. with only the Bz component; referred to as "inter- nal" fields). In the former case, the magnetic field plays a dominant role along the cloud surface and in the wake of the cloud where it reaches its highest strength (and the plasma β its lowest values; e.g. Mac Low et al. 1994; Jones et al. 1996). In the case of Bx, the magnetic field is trapped at the nose of the cloud, leading to a contin- uous increase of the magnetic pressure and field tension there (see upper panels in Fig. 2); in the case of By, the cloud expansion leads to the increase of magnetic pres- sure and field tension laterally to the cloud (see middle panels in Fig. 2). In the case of Bz (internal field), the magnetic field, being parallel to the cylindrical cloud, modifies only the total effective pressure of the plasma Anisotropic thermal conduction in SNRs 5 Fig. 2. -- As in Fig. 1 for the simulations with β0 = 4 and the magnetic field oriented along x (upper panels), y (middle panels), and z (lower panels). The figure shows the distribution in models either without (left half panels) or with (right half panels) thermal conduction and radiative losses. For runs NN-Bx4, TR-Bx4, NN-By4 and TR-By4, we plot the magnetic field lines; for runs NN-Bz4 and TR-Bz4, we include contours of log(B2/8π). 6 Orlando et al. (Jones et al. 1996); in the case of radiating shocks, the additional magnetic pressure may play a crucial role in the shocked cloud, preventing further compression of the cloud material (Fragile et al. 2005). 3.1.1. External magnetic fields In the case of predominantly external magnetic fields, Mac Low et al. (1994) and Jones et al. (1996) have shown that the hydrodynamic instabilities can be suppressed even in models neglecting the thermal conduction due to the tension of the magnetic field lines which maintain a more laminar flow around the cloud surface (see also Fragile et al. 2005): for a γ = 5/3 gas, the KH instabil- ities are suppressed if β < 2/M2, whereas RT instabili- ties are suppressed if β < (2/γ)(χ/M)2 (see also Chan- drasekhar 1961). However, for the parameters used in this paper (χ = 10 and M = 50), the magnetic field can- not suppress KH instabilities in any of our runs, whereas the RT instabilities are suppressed only in runs that lead to locally very strong field (β < 0.05). This can be seen in model NN-Bx4 (upper panels in Fig. 2), presenting a large field increase at the cloud boundary, compared to model NN (Fig. 1): in the latter case the growth of KH and RT instabilities at the cloud boundary is much more evident than in NN-Bx4. On the other hand, the hy- drodynamic instabilities are suppressed more efficiently in models including the thermal conduction (runs TR- Bx4 and TR-By4) even in cases with low field increase (for instance in our By case) as it is evident in Fig. 2 by comparing models NN-Bx4 and NN-By4 with models TR-Bx4 and TR-By4, respectively. The thermal exchanges between the cloud and the sur- rounding medium strongly depend on the initial field ori- entation. Fig. 3 shows the heat flux and magnetic field strength distributions in the (x, y) plane in runs TR- Bx4, TR-By4, and TR-Bz4, at time t = 2 τcc. In our Bx case (upper panels in Fig. 2), the magnetic field lines gradually envelope the cloud, reducing the heat conduc- tion through the cloud surface (see left panels in Fig. 3): thermal exchanges between the cloud and the surround- ing medium are channelled through small regions located at the side of the cloud. The cloud expansion and evap- oration are strongly limited by the confining effect of the magnetic field (cf. the unmagnetized case TR in Fig. 1 with model TR-Bx4 in Fig. 2) that becomes up to 30 times stronger just outside the cloud than inside it (see, also, the lower left panel in Fig. 3). The consequent ther- mal insulation induces the radiative cooling and conden- sation of the plasma into the cloud during the phase of cloud compression (t < τcc). At the end of this phase, the cloud material has temperature T ≈ 105 K and den- sity nH ≈ 10 cm−3 where primary and reverse shocks transmitted into the cloud are colliding; for these values of T and nH, the Field length scale (Begelman & McKee 1990) derived from the ratio of cooling timescale over conduction timescale (see Paper I for details) is l ≈ 106 T 2 nH ≈ 3.2 × 10−4 pc . (9) The radiative cooling dominates over the effects of the thermal conduction in cold and dense regions with di- mensions larger than l. At variance with our unmagne- tized case TR, therefore, thermal instabilities develop in run TR-Bx4. One of this cold and dense structures is ev- ident in Fig. 2 (upper panels) and is located at the cloud boundary near the nose of the cloud (at x ≈ 0.4 pc and y ≈ 3.0 pc) at t = 3 τcc. In the By case, the initial field direction is mostly main- tained in the cloud core during the evolution, allowing efficient thermal exchange between the core and the hot medium upwind of the cloud (see center panels in Fig. 3): the core is gradually heated and evaporates in few dy- namical timescales. This is illustrated by run TR-By4 in Fig. 2. On the other hand, the cloud is thermally in- sulated laterally where the magnetic field lines prevent thermal exchange between the cloud and the surrounding medium. Also, a strong magnetic field component along the x axis develops in the wake of the cloud and inhibits thermal conduction with the medium downwind of the cloud. The thermal insulation at the side of the cloud de- termines the growth of thermal instabilities where shocks transmitted into the cloud collide (see middle panels in Fig. 2). In both external field configurations, elongated struc- tures of strong field concentration are produced on the axis downwind of the cloud due to the focalization of the magnetized fluid flows there (see upper and middle panels of Fig. 2, and lower panels in Fig. 3). These fil- amentary structures, identified as "flux ropes" by Mac Low et al. (1994), are formed by magnetic field lines stretched around the cloud shape and do not carry a significant amount of cloud material (as shown by the tracer Ccl) although the plasma there moves with the cloud (see also Gregori et al. 2000). 3.1.2. Internal magnetic fields Predominantly internal magnetic fields strongly sup- press the heat conduction, providing an efficient ther- mal insulation of the cloud material (see right panels in Fig. 3). In the realistic configuration of an elongated cloud with finite length L along the z axis, some heat would be conducted along the magnetic field lines. The characteristic timescales for the conduction along mag- netic field lines is (see Paper I) τcond ≈ 2.6 × 10−9 nHL2 T 5/2 . (10) We estimate that the cloud would thermalize in τcond > 3.5 τcc (i.e. the physical time covered by our simula- tions), if the length scale of the cloud along the z axis is L > 3 pc. In this case, hydrodynamic instabilities de- velop at the cloud boundary, being both the magnetic field and the thermal conduction not able to suppress them. The growth of these instabilities is clearly seen in Fig. 2 (lower panels). The combined effect of hy- drodynamic instabilities and shocks transmitted into the cloud leads to unstable high-density regions at the cloud boundaries that trigger the development of thermal in- stabilities there (see lower panels in Fig. 2). However, as discussed by Fragile et al. (2005), internal magnetic field lines are expected to resist compression in the shocked cloud, thus reducing the cooling efficiency. In fact, in our run TR-Bz4, the cloud material is prevented from cooling below T ≈ 103 K. Since the thermal conduction does not play any significant role in the shock-cloud in- teraction, our Bz case leads to results similar to those Anisotropic thermal conduction in SNRs 7 Fig. 3. -- Heat flux (upper panels) and magnetic field strength (lower panels) distributions in the (x, y) plane in the simulations TR-Bx4 (left panels), TR-By4 (center), and TR-Bz4 (right), at time t = 2 τcc. The arrows in the upper panels describe the heat flux and scale linearly with respect to the reference value shown in the upper right corner of each panel. The scale of the magnetic field strength is linear and is given by the bar on the right, in units of 10 µG. The red contour encloses the cloud material. obtained by Fragile et al. (2005) and we do not discuss further this case. 3.2. Role of thermal conduction In this section, we study more quantitatively the effect of thermal conduction on the cloud evolution and, in par- ticular, on the cloud compression and on the magnetic field increase. To this end, we use the tracer defined in Sect. 2 to identify zones whose content is the original cloud material by more than 90%. Then, we define the cross-sectional area of cloud material, Acl(t), as the to- tal area in the (x, y) plane occupied by these zones. We define the cloud compression (or expansion) as Acl/Acl0, where Acl0 is the initial cross-sectional area. We also de- fine an average mass-weighted temperature of the cloud and an average magnetic field strength associated to the cloud as hT icl = hBicl = ZA(Ccl>0.9) ZA(Ccl>0.9) ZA(Ccl>0.9) ZA(Ccl>0.9) Ccl ρT da Ccl ρ da Ccl B da Ccl da (11) (12) where we integrate on zones with Ccl > 0.9. Note that our choice of considering cells with the value of the pas- sive tracer Ccl > 0.9 is arbitrary. To determine how sen- sitive the results are to this value and, in particular, to small changes in it, we derive our results also considering the values Ccl > 0.85 and Ccl > 0.95. In all the cases, we find that the results derived with the different thresholds show the same trend with differences lower than 10%. 8 Orlando et al. Fig. 4. -- Evolution of cloud compression (upper panels), of average temperature (middle panels), and of average magnetic field strength (lower panels) of the cloud for runs which neglect the thermal conduction and the radiation (dot-dashed lines; left panels), for runs which include the thermal conduction but neglect the radiation (solid; left panels), for runs which include the radiation but neglect the thermal conduction (dotted; right panels) and for runs which include both physical effects (dashed; right panels). The magnetized cases with β0 = 4 are marked with red (initial magnetic field along x), green (initial B along y) and blue (initial B along z) lines; the unmagnetized cases are marked with black lines. The light yellow regions mark the location of solutions which have thermodynamical characteristics in between the cases of maximum efficiency of the thermal conduction (models TN and TR) and the cases without thermal conduction (models NN and NR). By comparing the position of the magnetized models curves inside the yellow region, it is possible to quantitatively assess the degree of suppression of the effects of the thermal conduction by the magnetic fields. Fig. 4 shows the cloud compression, Acl/Acl0, the av- erage temperature of the cloud, hT icl, normalized to the post-shock temperature of the surrounding medium (Tpsh = 4.7 × 106 K), and the average magnetic field strength associated to the cloud, hBicl, normalized to the initial field strength (B0 = 1.31 µG, corresponding to β0 = 4) as a function of time for models neglecting thermal conduction and radiation (hereafter NNs mod- els), for models including conduction but neglecting radi- ation (TNs models), and for models including both con- duction and radiation (TRs models); we also include the results derived from the unmagnetized case NR with ra- diative cooling and without thermal conduction. The figure shows both the magnetized cases with β0 = 4 and the unmagnetized cases (see Table 2). In all the NNs models either with (NN-Bx4, NN-By4, and NN-Bz4) or without (NN) the magnetic field, the evolution of the cloud compression and of the average cloud temperature is roughly the same (see left panels in Fig. 4). The cloud is initially compressed over a timescale t ≈ τcc due to the ambient post-shock pressure; during this phase hT icl rapidly increases. After t ≈ τcc, the cloud partially reexpands, leading to a decrease of hT icl. In the last phase (t > 2.0 τcc), the cloud is compressed again by the interaction with the "Mach stem" formed during the reflection of the primary shock at the sym- metry axis, and hT icl increases; later Acl/Acl0 continues to decrease, because of the mixing of the cloud material Anisotropic thermal conduction in SNRs 9 with the ambient medium (see Sect. 3.3; see also Paper I), while hT icl stabilizes at ≈ 0.17 Tpsh. The field increase in the cloud material depends on the initial configuration of B (see lower left panel in Fig. 4). In the case of external fields (Bx and By components), B is mainly intensified due to stretching of field lines due to sheared motion. In the Bx case, the magnetic field undergoes the greatest increase and hBicl keeps increas- ing during the whole evolution. In fact the field is mainly intensified at the nose of the cloud where the background flow continues to stretch the field lines during the evolu- tion (see upper panels in Fig. 2). In the By case, the field increase occurs mainly at the side of the cloud where the field lines are stretched along the cloud surface. In the case of internal fields (Bz component), the field increase is due to squeezing of field lines through compression. hBicl, therefore, follows the changes in Acl/Acl0, since the field is locked within the cloud material. Thus the greatest field increase occurs at t ≈ τcc when the shocks transmitted into the cloud collide. The effects of thermal conduction are greatest in the unmagnetized model (TN) which can be considered an extreme limit case (see left panels in Fig. 4). During the first stage of evolution (t < 0.8 τcc), the cloud is heated efficiently by the thermal conduction and its av- erage temperature increases rapidly to ∼ 0.5 Tpsh. As a consequence, the pressure inside the cloud increases and the cloud reexpands earlier than in model NN. Af- terwards, the average cloud temperature, hT icl, keeps increasing up to ∼ 0.9 Tpsh at t = 3.5 τcc. In the case of predominantly external magnetic fields (models TN-Bx4 and TN-By4), the thermal conduction still plays a significant role in the cloud evolution, al- though its effects are not as large as in the unmagnetized case (TN). During the initial compression, the thermal conduction contributes to the cloud heating: the aver- age temperature of the cloud reaches values larger than in models neglecting the conduction (compare TNs with NNs models in the left panels of Fig. 4). This effect is greatest in the By case which is the configuration of field lines that allows the most efficient thermal exchange be- tween the cloud and the hot environment (see Sect. 3.1). At t = 3.5 τcc, hT icl in TNs models reaches values larger than in NNs models (≈ 0.21 Tpsh in the Bx case and ≈ 0.33 Tpsh in the By case). For internal magnetic fields, the thermal conduction plays no role in the evolution of the shocked cloud, being strongly ineffective due to B (see Sect. 3.1). As a consequence, the TN-Bz4 model leads to the same results as NNs models. In general, therefore, the effects of the thermal con- duction in the presence of an ambient magnetic field are reduced with respect to the corresponding unmag- netized case, but not entirely suppressed. This can be seen in Fig. 4, where we have marked in light yellow the region between the fully conductive unmagnetized case (TN) and the case without thermal conduction (NN). The magnetized TNs models are always within this re- gion, meaning that the effects of the thermal conduction are never as large as in the unmagnetized case (TN) but not completely suppressed as in the model NN. We also note that the thermal conduction influences indirectly the magnetic field increase. The main changes are in the By case and are due to the larger expansion of the cloud that reduces the increase of the field associ- ated to the cloud, being the field locked within the cloud material. In our unmagnetized case TR (including thermal con- duction and radiative cooling), the thermal conduction prevents the onset of thermal instabilities, and the evo- lution of the shocked cloud is the same as found in the TN model. At variance with our unmagnetized case TR, Fig. 4 shows that thermal instabilities develop in all our magnetized TRs runs, being the effects of thermal con- duction reduced by the magnetic field. The effects of radiative cooling are very strong for internal fields (our Bz case; see run TR-Bz4 in Fig. 4). In this case, the heat conduction is totally suppressed by the magnetic field and the evolution of Acl/Acl0 and of hT icl are the same as those found in the unmagnetized case with ra- diative cooling and without thermal conduction (model NR); at t = 3.5 τcc, run TR-Bz4 (and NR) shows the largest cloud compression (Acl/Acl0 ≈ 0.1) and the low- est cloud average temperature (hT icl ≈ 0.12 Tpsh). In the case of external fields (runs TR-Bx4 and TR-By4), the effects of heat conduction are reduced but not sup- pressed and the results are intermediate between those derived for runs NR and TR (i.e. within the light yellow region in the right panels in Fig. 4). The cooling effi- ciency is largely reduced in our By case (run TR-By4), namely that with the magnetic field configuration that allows the most effective thermal conduction. 3.3. Mass mixing and energy exchange We use the tracer to derive the cloud mass, Mcl, as the total mass in zones whose content is the original cloud material by more than 90%, Mcl = L ZA(Ccl>0.9) Ccl ρ da , (13) where L is the cloud length along the z axis, and the integral is done on zones with Ccl > 0.9. We investigate the mixing of cloud material with the ambient medium by defining the remaining cloud mass as Mcl/Mcl0, where Mcl0 is the initial cloud mass. The tracer allows us to investigate also the energy ex- change between the cloud and the surrounding medium; we derive the internal energy, Icl, and the kinetic energy, Kcl of the cloud as Ccl ρǫ da , Icl = L ZA(Ccl>0.9) 2 ZA(Ccl>0.9) L Kcl = Ccl ρu2 da , (14) (15) where again L is the cloud length along the z axis, and the integral is done on zones with Ccl > 0.9. We also define the total energy of the cloud as Ecl = Icl + Kcl . (16) Fig. 5 shows the evolution of the cloud mass, Mcl/Mcl0, for NNs, TNs, and TRs models; again we also include the unmagnetized case with radiative cooling and with- out thermal conduction (model NR). Both unmagnetized cases and magnetized cases with β0 = 4 are shown. In models without thermal conduction and radiation (NNs models), the hydrodynamic instabilities drive the mass 10 Orlando et al. Fig. 5. -- Presentation as in Fig. 4 for the evolution of the cloud mass (upper panels), of the internal energy of the cloud (middle panels), and of the kinetic energy of the cloud (lower panels). mixing of the cloud8. The mass loss rate of the cloud, mcl, increases significantly after 1.5 τcc (i.e. after the hydrodynamic instabilities have fully developed at the cloud boundary), with mcl ≈ 1.5 × 10−6Lpc M⊙ yr−1, where Lpc is the cloud length along the z axis in units of pc: ∼ 20% of the cloud mass is contained in mixed zones at t = 3.5 τcc. The only exception is run NN-Bx4 (∼ 15% of the cloud mass is in mixed zones at t = 3.5 τcc), being in this case RT instabilities partially suppressed by the magnetic field (compare run NN-Bx4 with runs NN-By4 and NN-Bz4 in Fig. 2). In TNs models with external magnetic fields (TN- Bx4 and TN-By4), the mass loss rate of the cloud is less efficient than in NNs models with mcl ≈ 6 × 10−7Lpc M⊙ yr−1 (∼ 10% of the cloud mass is in mixed zones at t = 3.5 τcc). In fact, in these cases the thermal 8 This is also true in our magnetized cases because, for the pa- rameters used in this paper (M = 50 and χ = 10), the hydrody- namic instabilities are partially suppressed by the magnetic field only in runs evolving to strong fields (see Sect. 3.1). conduction suppresses most of the hydrodynamic insta- bilities and the mass loss mainly comes from the cloud evaporation driven by the thermal conduction rather than from hydrodynamic ablation. Note that our unmag- netized TN model is an extreme limit case in which the hydrodynamic instabilities are totally suppressed by the thermal conduction which drives the cloud mixing; in this case, the mass loss rate is mcl ≈ 1.5 × 10−7Lpc M⊙ yr−1 (∼ 5% of the cloud mass is in mixed zones at t = 3.5 τcc). In magnetized TRs models, the onset of thermal in- stabilities increases the mass loss rate of the cloud with respect to the unmagnetized case ( mcl ranges between 1.5 × 10−6Lpc M⊙ yr−1 and 4 × 10−6Lpc M⊙ yr−1) due to the fragmentation of the cloud in dense and cold cloudlets. We expect, therefore, that the larger the amount of cloud mass mixed with the surrounding medium at the end of the evolution, the more limited the thermal exchange between the cloud and the hot ambi- ent medium (and, therefore, the greater the efficiency of radiative cooling). In fact, the upper right panel in Fig. 5 Anisotropic thermal conduction in SNRs 11 shows that the mass mixing has the greatest efficiency in run TR-Bz4 (i.e. in the case with the thermal conduction totally suppressed) which shows a mass loss rate of the cloud similar to that derived from the unmagnetized NR model. On the other hand, in runs TR-Bx4 and TR-By4, the mass mixing is intermediate between those derived with runs NR and TR. Fig. 5 also shows the evolution of internal (middle panels) and kinetic (lower panels) energy of the cloud, normalized to the initial total energy of the cloud, Ecl0. Among the magnetized cases considered, the greatest val- ues of Icl are reached in our By case which is the field configuration that allows the most efficient thermal ex- change between the cloud and the environment; the in- crease of Icl is due to the heat conducted to the shocked cloud. Also, the By case leads to the greatest values of Kcl because the cloud has a larger cross-sectional area (because of the larger cloud expansion due to the heat- ing driven by heat conduction; see upper panels in Fig. 4) and offers, therefore, a larger surface to the pressure of the shock front responsible of the cloud acceleration. 3.4. Role of the initial field strength In this section we explore the effects of the initial field strength on the mass mixing and energy exchange of the cloud. Fig. 6 shows the evolution of the cloud mass, Mcl/Mcl0 (upper panel), and of the total (internal plus kinetic) energy of the cloud, Ecl/Ecl0 (lower panel), for magnetized TRs models with different values of β0. We discuss here only the cases of predominantly external magnetic fields (Bx or By case) since no significant de- pendence on the initial field strength has been found in the case of predominantly internal magnetic fields (Bz case). Fig. 6 shows that the initial field strength plays a sig- nificant role in the Bx case. In particular, models with greater values of β0 show a more efficient mixing of the cloud material and a less rapid increase of the cloud en- ergy. As discussed in Sect. 3.3, in the Bx case, the rate of mass-loss from the cloud is mainly driven by abla- tion through the hydrodynamic instabilities (being the thermal conduction strongly suppressed by the magnetic field). On the other hand, in the case of external fields, the instabilities can be dumped by the magnetic field, de- pending on its strength (see Sect. 3.1.1). For instance, in the Bx case with β0 = 4, we found that the RT insta- bilities are mostly suppressed by the magnetic field (see upper panels in Fig. 2). In the Bx case with β0 = 100, instead the magnetic field is too weak to dump the hy- drodynamic instabilities over the timescales considered; these instabilities, in turn, lead to the formation of re- gions dominated by the radiative cooling, triggering the development of thermal instabilities. Both the hydrody- namic and the thermal instabilities determine the cloud mass mixing (which is higher for higher values of β0). In addition, the thermal instabilities reduce the increase of the cloud energy (which is less rapid for higher β0) due to significant radiative losses. In the By case, the initial field strength has a smaller influence on the dynamic and thermal evolution of the cloud than in the Bx case (see Fig. 6). In addition, at variance with the Bx case, models with greater values of β0 show a less efficient mixing of the cloud material and a more rapid increase of the cloud energy. In the By Fig. 6. -- Evolution of the cloud mass (upper panel) and of the total energy of the cloud (internal plus kinetic; lower panel) for runs including both the thermal conduction and the radiative cooling (TRs models). The figure shows the simulations with the magnetic field oriented along x (red lines) or y (green) and with β0 = 1 (dotted lines), 4 (solid), and 100 (dashed). case, in fact, the hydrodynamic instabilities responsible of the mass mixing are mainly suppressed by the thermal conduction rather than by the magnetic field as in the Bx case. As a consequence, the higher the value of β0, the more effective the thermal conduction in suppressing the instabilities and in heating the plasma, the less efficient the cloud mass mixing and the more rapid the increase of the cloud energy. 3.5. Effect of spatial resolution The effective resolution adopted in our simulations is ≈ 132 cells per cloud radius, a value above the resolu- tion requirements suggested by Klein et al. (1994) for non-radiative clouds. However, for radiative clouds, we expect that the details of the plasma radiative cooling depend on the numerical resolution: a higher resolution may lead to different peak density and hence influence the cooling efficiency of the gas, preventing further com- pression of the cloud. In the non-conducting regime, Fragile et al. (2005) found that the results generally con- verge for simulations with resolution larger than 100 cells per cloud radius ( < In the simula- tions presented here, the thermal conduction partially contrasts the radiative cooling in the case of external fields (Bx or By), alleviating the problem of numerical resolution (see also Paper I). ∼ 10% differences). 12 Orlando et al. the latters reaches the resolution limit toward the end of the simulations when the relevant physical processes are already at a late stage. 4. SUMMARY AND CONCLUSION We investigated the importance of magnetic-field- oriented thermal conduction in the interaction between an isolated elongated dense cloud and an interstellar shock-wave of an evolved SNR shell through numerical MHD simulations. To our knowledge, these simulations represent the first attempt to model the shock-cloud in- teraction that simultaneously considers magnetic fields, radiative cooling, and anisotropic thermal conduction. Our findings lead to several conclusions: 1. In general, we found that the effects of thermal con- duction on the evolution of the shocked cloud are reduced in the presence of an ambient magnetic field with respect to the unmagnetized cases in- vestigated in Paper I. The efficiency of anisotropic thermal conduction strongly depends on the ini- tial magnetic field orientation and configuration. This efficiency is the largest when the initial B is aligned with the direction of propagation of the shock front, and is the smallest when B is aligned with the cylindrical cloud, namely when the heat conduction is completely suppressed by the mag- netic field. 2. We found that the hydrodynamic instabilities are suppressed efficiently by the anisotropic thermal conduction when the initial magnetic field is per- pendicular to the cylindrical cloud (a configuration referred to as "external fields"). On the contrary, in the case of B parallel to the cylindrical axis of the cloud (i.e. when the field has component only along the z axis - internal field), hydrodynamic instabili- ties develop at the cloud boundary. We found that, for the parameters of the simulations chosen, the magnetic tension is unable to suppress alone the hydrodynamic instabilities. 3. As for thermal instabilities, we found that, depend- ing on the magnetic field orientation, the heat flux contributes to the heating of some parts of the cloud, reducing the efficiency of radiative cooling there, and preventing any thermal instability. 4. The mass loss of the cloud due to mixing with the surrounding medium is mainly driven by hy- drodynamic instabilities; in the case of external fields (initial B perpendicular to the cylindrical cloud) the anisotropic thermal conduction reduces the mass mixing of the cloud. In any case, the mass loss rate is larger than that in the corresponding unmagnetized case ( mcl ≈ 1.5×10−7Lpc M⊙ yr−1, i.e. ∼ 5% of the cloud mass is in mixed zones at t = 3.5 τcc), but can get very high when the thermal conduction is completely suppressed ( mcl ≈ 4 × 10−6Lpc M⊙ yr−1, i.e. ∼ 45% of the cloud mass is in mixed zones at t = 3.5 τcc). 5. The thermal conduction mostly rules the en- ergy exchange between the cloud and surrounding Fig. 7. -- Presentation as in Fig. 6 for runs TR-Bz4 (solid lines), TR-Bz4-hr (dashed), and TR-Bz4-hr2 (dotted). In order to check if our adopted resolution is sufficient to capture the basic cloud evolution over the time inter- val considered, we compare three simulations (TR-Bz4, TR-Bz4-hr, and TR-Bz4-hr2) with different spatial reso- lution (132, 264, and 528 zones per cloud radius, respec- tively) for the Bz case with β = 4, namely one of the cases in which the growth of hydrodynamic and thermal insta- bilities is most prominent and the effect of thermal con- duction (contrasting the development of hydrodynamic and thermal instabilities) is negligible. Since this case is one of the most demanding for resolution, it can be considered a worst case comparison of convergence. Figure 7 compares the evolution of the cloud mass, Mcl/Mcl0, and of the total energy of the cloud, Ecl/Ecl0, for the three simulations TR-Bz4, TR-Bz4-hr, and TR- Bz4-hr2. In general, we find that the results obtained with the three simulations agree quite well in their quali- tative behavior, showing differences < ∼ 10%. In runs TR- Bz4-hr and TR-Bz4-hr2, the remaining cloud mass and the total energy of the cloud are, in general, systemati- cally higher than in run TR-Bz4. The larger mass mixing in TR-Bz4 is driven by the higher diffusion of the low- resolution grid down to the very small structures which tend to smear out concentrated density peaks, promot- ing mass mixing. The slightly lower energy of the cloud in TR-Bz4 is a consequence of the larger mass mixing derived in this run with respect to TR-Bz4-hr and TR- Bz4-hr2. Note that, in runs showing the onset of ther- mal instabilities (i.e NRs and TRs models), the size of Anisotropic thermal conduction in SNRs 13 medium. The exchange is favored when the mag- netic field configuration is such that the conductive flow is not suppressed (i.e. external field configu- rations, Bx and By cases), but it is never as high as in the absence of magnetic field. In the By case, the cloud core is efficiently heated and evaporates in few dynamical timescales. 6. In general, the initial magnetic field strength has a small influence on the dynamic and thermal evolu- tion of the shocked cloud for the ranges of values explored in this paper (namely 0.26 µG ≤ B ≤ 2.63 µG). It is worth noting that some details of our simulations depend on the choice of the model parameters. For in- stance, the onset of thermal instabilities or the evapora- tion of the whole cloud depends on the initial shock Mach number, and on the density and dimensions of the cloud. The cases that we present here (i.e. M = 50, χ = 10, and different configurations of B) are representative of a regime in which both the thermal conduction and the radiative cooling play an important role in the evolution of the shocked cloud. Nevertheless, our analysis proves that anisotropic thermal conduction can not be neglected in investigations of the evolution of shocked interstellar clouds. In our simulations, we consider laminar thermal con- duction, although regions of strong turbulence of differ- ent strength and extent develop in the system (for in- stance, at the shear layers along the cloud boundary or at the vortex sheets in the cloud wake). In fact, the turbulence in these regions may have a significant ef- fect on thermal conduction, leading to significant devi- ations of thermal conductivity from its laminar values (e.g. Narayan & Medvedev 2001; Lazarian 2006); in some cases, the turbulence may enhance the heat transfer, ex- ceeding the classical Spitzer value (Lazarian 2006). As a result, thermal conduction may be not only anisotropic (in the presence of the magnetic field) but also "inhomo- geneous" due to the presence of turbulence. However, even modeling accurately the turbulent thermal conduc- tivity, we do not expect significant changes in the results of our Bz case, being the thermal conduction strongly ineffective in the whole spatial domain; in the remaining cases (Bx and By), our modeled thermal conductivity could be underestimated in regions of strong turbulence, affecting some details of the simulations but not the main conclusion of the paper that, in general, anisotropic ther- mal conduction can play an important role in the evolu- tion of the shocked cloud. Note also that the field configurations studied in this work are highly idealized. More realistic fields are ex- pected to have more complex topologies and, often, the field can be tangled and chaotic. In the latter case, the thermal conduction will approach isotropy, whereas the effect of MHD turbulence is expected to partially sup- press the heat transfer within a factor ∼ 5 below the classical Spitzer estimate9 (Narayan & Medvedev 2001; Lazarian 2006). The shock-cloud collision in the presence of an organized ambient magnetic field, discussed here, 9 As already discussed, the MHD turbulence can even enhance the heat transfer in some cases (see Lazarian 2006). and that in the absence of magnetic field can be con- sidered as extreme cases: the former leading to highly anisotropic thermal conduction, the latter to the clas- sical Spitzer thermal conduction. The case of chaotic magnetic field is expected to fall in between these two. Our simulations were carried out in 2.5D Cartesian geometry, implying that the modeled clouds are elon- gated along the z axis. This choice is expected to affect some details of the simulations but not our main con- clusions. Adopting a 3D Cartesian geometry and model- ing a spherical cloud, the highly symmetric shock trans- mitted into the cloud converging on the symmetry axis would lead to compression stronger than those found in our 2.5D simulations, enhancing the radiative cooling. Also, 3D simulations would provide an additional degree of freedom for hydrodynamic instabilities, increasing the mass loss rate of the cloud in the cases in which the mass mixing of cloud material is driven by instabilities. Note that, for a spherical cloud, our Bx and Bz cases no longer differ. Finally we assume, in our simulations, that the cloud and the ambient material have the same composition, implying that microscopic mass mixing due to shear in- stabilities would be irrelevant. In a more realistic condi- tion, a cold dense cloud may have a different composition from the hot ambient flow and the degree of microscopic mixing may translate into different spectral signatures of the system. In this case, species diffusion could also be important, along with thermal conduction, to determine the degree of microscopic mixing of the materials and, consequently, one would have to ask about the typical values of the Lewis number (i.e. the ratio of thermal diffusivity to mass diffusivity) in the system. It is worth emphasizing that the quantitative results of our simulations depend on the physical parameters of the model (shock Mach number, density contrast and dimension of the cloud, etc.) as well as on the basic assumptions of the model (geometry of the cloud, ge- ometry of the ambient magnetic field, laminar thermal conduction, composition of the cloud and of the ambient medium, etc.). Nevertheless, our results undoubtedly show that the magnetic-field-oriented thermal conduc- tion can play an important role in the evolution of the shock-cloud interaction (which depends on the magnetic field orientation and configuration) and, in particular, in the mass and energy exchange between the cloud and the hot surrounding medium. We conclude, therefore, that a self-consistent and quantitative description of the interaction between magnetized shock-waves and inter- stellar gas clouds should include the effects of thermal conduction. The results presented here are interesting for the study of middle-aged SNR shells expanding into a magnetized ISM and whose morphology is affected by ISM inhomo- geneities (for instance, G272.2-3.2, e.g. Egger et al. 1996; Cygnus Loop, e.g. Patnaude et al. 2002; Vela SNR, e.g. Miceli et al. 2005). It will be further interesting to ex- tend the present study, by modeling the shock-cloud in- teraction in 3D with radiative cooling, anisotropic ther- mal conduction, and magnetic field included and, even, considering detailed comparisons of model results with observations. 14 Orlando et al. The authors thank Timur Linde for his help with the MHD portion of FLASH and the referee for constructive and helpful criticism. The software used in this work was in part developed by the DOE-supported ASC / Al- liance Center for Astrophysical Thermonuclear Flashes at the University of Chicago, using modules for ther- mal conduction and optically thin radiation built at the Osservatorio Astronomico di Palermo. Most of the simulations have been executed at CINECA (Bologna, Italy) in the framework of the INAF-CINECA agreement on "High Performance Computing resources for Astron- omy and Astrophysics". This work makes use of results produced by the PI2S2 Project managed by the Con- sorzio COMETA, a project co-funded by the Italian Min- istry of University and Research (MIUR) within the Pi- ano Operativo Nazionale "Ricerca Scientifica, Sviluppo Tecnologico, Alta Formazione" (PON 2000-2006); more information is available at http://www.pi2s2.it and http://www.consorzio-cometa.it. This work was sup- ported in part by Istituto Nazionale di Astrofisica. Balbus, S. A. 1986, ApJ, 304, 787 Begelman, M. C., & McKee, C. F. 1990, ApJ, 358, 375 Bocchino, F., Maggio, A., Sciortino, S., & Raymond, J. 2000, A&A, MacNeice, P., Olson, K. M., Mobarry, C., de Fainchtein, R., & Packer, C. 2000, Comp. Phys. Comm., 126, 330 Mellema, G., Kurk, J. D., & Rottgering, H. J. A. 2002, A&A, 395, 359, 316 L13 REFERENCES Borkowski, K. J., Shull, J. M., & McKee, C. F. 1989, ApJ, 336, 979 Chandrasekhar, S. 1961, Hydrodynamic and hydromagnetic stability (International Series of Monographs on Physics, Oxford: Clarendon, 1961) Cowie, L. L., & McKee, C. F. 1977, ApJ, 211, 135 Egger, R., Greiner, 1996, in Roentgenstrahlung from the Universe, MPE Report 263, 247 -- 248 J., & Aschenbach, B. Einfeldt, B. 1988, SIAM J. Numer.Anal, 25, 357 Fadeyev, Y. A., Le Coroller, H., & Gillet, D. 2002, A&A, 392, 735 Fragile, P. C., Anninos, P., Gustafson, K., & Murray, S. D. 2005, ApJ, 619, 327 Mewe, R., Gronenschild, E. H. B. M., & van den Oord, G. H. J. 1985, A&AS, 62, 197 Miceli, M., Bocchino, F., Maggio, A., & Reale, F. 2005, A&A, 442, 513 Miceli, M., Reale, F., Orlando, S., & Bocchino, F. 2006, A&A, 458, 213 Nakamura, F., McKee, C. F., Klein, R. I., & Fisher, R. T. 2006, ApJS, 164, 477 Narayan, R., & Medvedev, M. V. 2001, ApJ, 562, L129 Orlando, S. Bocchino, F., Peres, G., Reale, F., Plewa, T., & Rosner, R. 2006, A&A, 457, 545 Orlando, S., Peres, G., Reale, F., Bocchino, F., Rosner, R., Plewa, Fragile, P. C., Murray, S. D., Anninos, P., & van Breugel, W. 2004, T., & Siegel, A. 2005, A&A, 444, 505 ApJ, 604, 74 Pagano, P., Reale, F., Orlando, S., & Peres, G. 2007, A&A, 464, Fryxell, B. et al. 2000, ApJS, 131, 273 Giuliani, J. L. 1984, ApJ, 277, 605 Gregori, G., Miniati, F., Ryu, D., & Jones, T. W. 2000, ApJ, 543, 775 Jones, T. W., Ryu, D., & Tregillis, I. L. 1996, ApJ, 473, 365 Kaastra, J. S., & Mewe, R. 2000, in Atomic Data Needs for X-ray Astronomy, p. 161 Klein, R. I., McKee, C. F., & Colella, P. 1994, ApJ, 420, 213 Lazarian, A. 2006, ApJ, 645, L25 Lohner, R. 1987, Comp. Meth. Appl. Mech. Eng., 61, 323 Mac Low, M., McKee, C. F., Klein, R. I., Stone, J. M., & Norman, M. L. 1994, ApJ, 433, 757 Mac Low, M.-M., & Klessen, R. S. 2004, Reviews of Modern Physics, 76, 125 753 Patnaude, D. J., Fesen, R. A., Raymond, J. C., Levenson, N. A., Graham, J. R., & Wallace, D. J. 2002, AJ, 124, 2118 Raymond, J. C., & Smith, B. W. 1977, ApJS, 35, 419 Spitzer, L. 1962, Physics of Fully Ionized Gases (New York: Interscience, 1962) Zel'dovich, Y. B., & Raizer, Y. P. 1966, Physics of Shock Waves and High-Temperature Hydrodynamic Phenomena (New York: Academic Press, 1966)
astro-ph/0105473
1
0105
2001-05-28T04:52:23
Stable Operation of a 300-m Laser Interferometer with Sufficient Sensitivity to Detect Gravitational-Wave Events within our Galaxy
[ "astro-ph", "gr-qc" ]
TAMA300, an interferometric gravitational-wave detector with 300-m baseline length, has been developed and operated with sufficient sensitivity to detect gravitational-wave events within our galaxy and sufficient stability for observations; the interferometer was operated for over 10 hours stably and continuously. With a strain-equivalent noise level of $h\sim 5 \times 10^{-21} /\sqrt{\rm Hz}$, a signal-to-noise ratio (SNR) of 30 is expected for gravitational waves generated by a coalescence of 1.4 $M_\odot$-1.4 $M_\odot$ binary neutron stars at 10 kpc distance. %In addition, almost all noise sources which limit the sensitivity and which %disturb the stable operation have been identified. We evaluated the stability of the detector sensitivity with a 2-week data-taking run, collecting 160 hours of data to be analyzed in the search for gravitational waves.
astro-ph
astro-ph
Stable Operation of a 300-m Laser Interferometer with Sufficient Sensitivity to Detect Gravitational-Wave Events within our Galaxy Masaki Ando,1 , ∗ Ko ji Arai,2 Ryutaro Takahashi,2 Gerhard Heinzel,2 Seiji Kawamura,2 Daisuke Tatsumi,2 Nobuyuki Kanda,3 Hideyuki Tagoshi,4 Akito Araya,5 Hideki Asada,6 Youich Aso,1 Mark A. Barton,7 Masa-Katsu Fujimoto,2 Mitsuhiro Fukushima,2 Toshifumi Futamase,8 Kazuhiro Hayama,9 Gen’ichi Horikoshi,10 , † Hideki Ishizuka,7 Norihiko Kamikubota,10 Keita Kawabe,1 Nobuki Kawashima,11 Yoshinori Kobayashi,1 Yasufumi Ko jima,12 Kazuhiro Kondo,7 Yoshihide Kozai,2 Kazuaki Kuroda,7 Namio Matsuda,13 Norikatsu Mio,14 Kazuyuki Miura,3 Osamu Miyakawa,7 Shoken M. Miyama,2 Shinji Miyoki,7 Shigenori Moriwaki,14 Mitsuru Musha,15 Shigeo Nagano,16 Ken’ichi Nakagawa,15 Takashi Nakamura,17 Ken-ichi Nakao,18 Kenji Numata,1 Yujiro Ogawa,10 Masatake Ohashi,7 Naoko Ohishi,2 Satoshi Okutomi,7 Ken-ichi Oohara,19 Shigemi Otsuka,1 Yoshio Saito,10 Misao Sasaki,4 Shuichi Sato,7 Atsushi Sekiya,1 Masaru Shibata,4 Kentaro Somiya,14 Toshikazu Suzuki,10 Akiteru Takamori,1 Takahiro Tanaka,17 Shinsuke Taniguchi,1 Souichi Telada,20 Kuniharu Tochikubo,1 Takayuki Tomaru,7 Kimio Tsubono,1 Nobuhiro Tsuda,21 Takashi Uchiyama,10 Akitoshi Ueda,2 Ken-ichi Ueda,15 Koichi Waseda,2 Yuko Watanabe,3 Hiromi Yakura,3 Kazuhiro Yamamoto,1 and Toshitaka Yamazaki2 (the TAMA collaboration) 1Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan 2National Astronomical Observatory of Japan, Mitaka, Tokyo 181-8588, Japan 3Department of Physics, Miyagi University of Education, Aoba Aramaki, Sendai 980-0845, Japan 4Department of Earth and Space Science, Osaka University, Toyonaka, Osaka 560-0043, Japan 5Earthquake Research Institute, University of Tokyo, Bunkyo-ku, Tokyo 113-0032, Japan 6Faculty of Science and Technology, Hirosaki University, Hirosaki, Aomori 036-8561, Japan 7 Institute for Cosmic Ray Research, University of Tokyo, Kashiwa, Chiba 277-8582, Japan 8Astronomical Institute, Tohoku University, Sendai, Miyagi 980-8578, Japan 9Department of Astronomy, University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan 10High Energy Accelerator Research Organization, Tsukuba, Ibaragi 305-0801, Japan 11Department of Physics, Kinki University, Higashi-Osaka, Osaka 577-8502, Japan 12Department of Physics, Hiroshima University, Higashi-Hiroshima, Hiroshima 739-8526, Japan 13Department of Materials Science and Engineering, Tokyo Denki University, Chiyoda-ku, Tokyo 101-8457, Japan 14Department of Advanced Materials Science, University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan 15 Institute for Laser Science, University of Electro-Communications, Chofugaoka, Chofu, Tokyo 182-8585, Japan 16Max-Planck-Institut fur Quantenoptik, Cal linstrasse 38, D-30167 Hannover, Germany 17Yukawa Institute for Theoretical Physics, Kyoto University, Kyoto 606-8502, Japan 18Department of Physics, Osaka City University, Sumiyoshi-ku, Osaka, Osaka 558-8585, Japan 19Department of Physics, Niigata University, Niigata, Niigata 950-2102, Japan 20National Research Laboratory of Metrology, Tsukuba, Ibaragi 305-8563, Japan 21Precision Engineering Division, Tokai University, Hiratsuka, Kanagawa 259-1292, Japan (Dated: February 1, 2008) TAMA300, an interferometric gravitational-wave detector with 300-m baseline length, has been developed and operated with sufficient sensitivity to detect gravitational-wave events within our galaxy and sufficient stability for observations; the interferometer was operated for over 10 hours stably and continuously. With a strain-equivalent noise level of h ∼ 5× 10−21 /√Hz, a signal-to-noise ratio (SNR) of 30 is expected for gravitational waves generated by a coalescence of 1.4 M⊙ -1.4 M⊙ binary neutron stars at 10 kpc distance. We evaluated the stability of the detector sensitivity with a 2-week data-taking run, collecting 160 hours of data to be analyzed in the search for gravitational waves. PACS numbers: 04.80.Nn, 07.60.Ly, 95.55.Ym Introduction. — The direct observation of gravita- tional waves (GW) is expected to reveal new aspects of the universe [1]. Since GWs are emitted by the coher- ent bulk motion of matter, and are hardly absorbed or scattered, they carry different information from that of electromagnetic waves. However, no GW has yet been detected directly because of its weakness. In order to cre- ate a new field of GW astronomy, several groups around the world are developing laser interferometric GW de- tectors. Compared with resonant-type GW detectors [2], interferometric detectors have an advantage in that they can observe the waveform of a GW, which would contain astronomical information. Interferometric GW detectors are based on a Michelson interferometer. The quadrupole nature of a GW causes differential changes in the arm lengths of the Michelson interferometer, which are detected as changes in the in- terference fringe. Interferometric detectors have been in- vestigated with many table-top [3] and prototype [4] ex- periments to evaluate the principle of GW detection and their potential sensitivity to GWs. With the knowledge obtained from these experimental interferometers, sev- eral GW detectors with baseline lengths of 300 m to 4 km are under construction: LIGO [5] in U.S.A., VIRGO [6] and GEO [7] in Europe, and TAMA [8] in Japan. In these detectors, both high sensitivity and high stability are required because the GW signals are expected to be extremely small and rare. TAMA is a Japanese pro ject to construct and operate an interferometric GW detector with a 300-m baseline length at the Mitaka campus of the National Astronom- ical Observatory in Tokyo (35◦40′N, 139◦32′E). In this article, we report on an important achievement in inter- ferometric detectors: the TAMA detector was operated with sufficient sensitivity and stability to observe GW events at the center of our galaxy. The interferometer was operated stably and continuously over several hours in typical cases, and over 10 hours in the best cases. The noise-equivalent sensitivity was h ∼ 5 × 10−21 /√Hz at the floor level (700 Hz to 1.5 kHz). With this stability of the sensitivity, TAMA has the ability to detect GW events throughtout much of our galaxy: chirp signals from the coalescence of binary neutron stars or binary MACHO black holes [9], and burst signals from super- nova explosions. Detector configuration. — In the interferometer, called TAMA300, the arms of the Michelson interferometer are replaced by 300 m Fabry-Perot arm cavities to en- hance the sensitivity to GWs (Fig. 1). The arm cavities have a finesse of around 500, and a cutoff frequency of about 500 Hz; the light is stored in the cavities for about 0.3 msec. Since a high-power and stable laser is required as a light source, we use an LD-pumped Nd:YAG laser with an output power of 10 W [10]. In addition, a mode cleaner is inserted between the laser source and the main interferometer to reject higher-mode beams and to stabi- lize the laser frequency. The mode cleaner of TAMA300 is an independently-suspended triangular ring cavity with a length of 9.75 m [11]. Electro-optic modulators (EOM) for phase modulation (for mode cleaner and the main interferometer control) are placed in front of the mode cleaner. Thus, the wave-front distortion by the EOM are rejected by the mode cleaner before entering into the main interferometer. The mirrors of the main interferometer are made of fused silica. Each mirror has a diameter of 100 mm, and a thickness of 60 mm. The mirrors are coated by an IBS (ion-beam sputtering) machine to realize low-optical-loss surfaces [12]. The mirrors of the main interferometer and the mode 2 Main interferometer Michelson interferometer with 300-m Fabry-Perot arm cavities Arm cavity alignment control Differential motion control (dL-) Arm cavity alignment control Fringe control (dl-) Input beam axis control To input beam steering mirror Gravitational wave signal (dL-) Photo Detector Quadrant Photo Detector Mixer Injection-locked Nd:YAG 10-W laser 10-m ring-type Mode cleaner MC alignment control BS orientation control Laser frequency stabilization MC mass loop Feed around loop Common motion control (dL+) FIG. 1: Optical and control design of the TAMA300 interfer- ometer. TAMA300 is a Fabry-Perot-Michelson interferometer with a baseline length of 300 m. A triangular ring cavity is inserted between the main interferometer and the laser source as a mode cleaner. The control system is designed to realize high sensitivity and stability of the detector at the same time. cleaner are isolated from seismic motion by over 165 dB (at 150 Hz) with three-stage stacks [13] and double- pendulum suspension systems [14]. The suspension points are fixed to motorized stages, which are used for an initial adjustment of the mirror orientations. The fine position and orientation of each mirror is controlled with coil-magnet actuators; small permanent magnets are at- tached to the mirror. The interferometer is housed in a vacuum system com- prising eight chambers connected with beam tubes with a diameter of 400 mm. With surface processing, called ECB (Electro-Chemical Buffing), a vacuum pressure of less than 10−6 Pa is achieved without baking [15]. The control system is designed to realize high sensitiv- ity and stability at the same time. It consists of three parts: a length control system to keep the interferometer at its operational point, an alignment control system to realize short-term (∼ 1 minutes) stability and high sensi- tivity, and a beam-axis drift control system for long-term (∼ a few hours) stable operation. A frontal modulation scheme [16] is used for the length control; 15.235 MHz phase modulation is used for signal extraction. The dif- ferential motion signal of the arm cavities (δL− ) is fed back to the front mirrors with a bandwidth of 1 kHz. The common motion signal (δL+ ) is fed back to the mode cleaner and the laser source to stabilize the laser fre- quency. The motion in the Michelson interferometer part (δ l− ) is fed back to the beam splitter. An alignment control system is necessary for stable and sensitive operation because angular fluctuations (about several µrad) of the suspended mirrors excited by seis- mic motion make the interferometer unstable. The con- trol signals are extracted by a wave-front-sensing scheme [17], and fed back to each mirror; the angular motions are suppressed by over 40 dB to 10−8 rad in root-mean- square. Low-frequency drift control of the laser beam axis plays an important role in maintaining long-term operation. The beam axes are controlled with 300 m optical levers; the beam positions of the light transmitted through the arm cavities are monitored with quadrant photo detec- tors, and are fed back to the input steering mirror and the beam splitter of the main interferometer. The data-acquisition system comprises a high- frequency part for the main signals and a low-frequency part for detector diagnostics. The main output signals of the interferometer are recorded with high-frequency A/D converters (2× 104 samples/sec, 16 bit) after passing through whitening filters and 5 kHz anti-aliasing low-pass filters. Seven channel signals are recorded together with a timing signal, which provides a GPS-derived coordi- nated universal time (UTC) within an accuracy of 1 µsec. Along with the high-frequency system, 88 channels of monitoring signals are collected with a low-frequency data-acquisition system for interferometer diagnosis. Detector noise level. — Figure 2 shows the typical noise level of TAMA300 (black curve). The displacement noise level of the interferometer is 1.5 × 10−18 m/√Hz, which corresponds to 5 × 10−21 /√Hz in strain. Almost all of the noise sources which limit the interferometer noise level have been identified. The gray curve in Fig. 2 represents the total contribution of the identified noise sources: seismic motion (∼ 30 Hz), alignment-control noise (30 Hz ∼ 300 Hz), Michelson phase-detection noise (300 Hz ∼ 3 kHz), and the laser frequency noise (3 kHz ∼). The seismic noise and the laser frequency noise are esti- mated to satisfy the design requirements in the observa- tion band (around 300 Hz). Though the alignment control system is indispensable for the stable operation of the interferometer, this sys- tem can introduce excess noise to the interferometer [18]. The noise in the alignment control error signal causes displacement noise by coupling with mis-centering of the beam on a mirror, and efficiency asymmetries of the coil- magnet actuators on a mirror. In order to reduce this noise, the actuator balances are adjusted so that the ro- tational center is at the beam spot on a mirror. In ad- dition, the beam position on each mirror is controlled to be still by the beam-axis control system. The noise floor level is limited by the phase-detection noise of the Michelson part of the interferometer. Phase changes in the light caused by GW are amplified in the arm cavities, and detected by the Michelson interferom- eter. In order to realize the designed detector sensitivity, the noise level of the Michelson part of the interferome- ter should be limited only by the shot noise in the ob- servation band. However, in TAMA, as well as by the shot noise, it is currently limited by the scattered light 10–12 10–14 10–16 10–18 10–20 ] 2 / 1 z H / 1 [ e s i o n n i a r t S 10–22 101 3 Interferometer noise level S e i s m i c n o i s e A lig n m c e o n n t tr ol n ois e Michelson phase noise 102 103 Frequency [Hz] e o i s y n c n e r q e e u L a s f r 104 FIG. 2: Noise level of the TAMA300 interferometer (black curve) and the total contribution of identified noise sources (gray curve). The floor level is 5 × 10−21 /√Hz in strain. The thick curves represent the contribution of the noise sources. noise caused by the anti-reflection coating of the mode- matching telescope (MMT in Fig. 1) between the main interferometer and the mode cleaner. To realize the de- signed detector sensitivity which is purely limited by the shot noise, the scattering noise should be removed. Stability of operation. — With the total system de- scribed above, we performed a 2-week observation run from August 21 to September 3, 2000. Figure 3 shows the operational state during the observation; the gray and black boxes represent the time when the interferometer was operated and when the data were taken, respectively. The time is shown both in UTC and in Japan standard time (JST). The interferometer was operated for over 160 hours, 94.8% of the total data-taking-run time. (Peri- ods of continuous lock shorter than 10 minutes are not included.) We operated the interferometer mainly dur- ing the night for the efficient collection of high-quality data. During the observation, the noise level was de- graded slightly from the Fig. 2 level, because of electronic noises by many cables connected to the interferometer for data-taking. The longest continuous locking time was over 12 hours (several hours in typical cases); the main cause of loss of lock of the interferometer was large seis- mic disturbances, including earthquakes (closed circles), and rather large drifts of δL+ (triangle marks). From a coincidence analysis of signals from seismometers placed at the center and end rooms, we found that the inter- ferometer was knocked out of lock by accelerations of 12 mgal (1 ∼ 10 Hz frequency range). The interferometer was also knocked out of lock by large drift of δL+ ; a drift of over 120 µm could cause saturation in the feed- back loop. The other causes of loss of lock are thought to be local seismic disturbances caused by human activities, spikes due to instability of the laser source, and so on. The interferometer noise level was stable thanks to the alignment and drift control systems; the drift of the noise 15:00 18:00 21:00 00:00 03:00 06:00 09:00 Time in JST 09:00 12:00 Aug. 21 22 Operated (over 10min) High–freq. data taking Seismic disturbance Cavity length drift 23 24 25 26 27 28 29 30 31 Sept. 02 03 00:00 03:00 Time in UTC 06:00 09:00 12:00 15:00 18:00 21:00 24:00 FIG. 3: Operation status of the TAMA interferometer during a data-taking run performed from August 21 to September 3 in the year 2000. The interferometer was operated stably for over 160 hours, 94.8% of the total data-taking-run time. 10(cid:13) 9(cid:13) 8(cid:13) 7(cid:13) 6(cid:13) 5(cid:13) 4(cid:13) 3(cid:13) 2(cid:13) 1(cid:13) ](cid:13) (cid:13)o (cid:13)M [ s s a m l a t o T 0(cid:13) 0 1 0(cid:13) 5 0(cid:13) 3 0(cid:13) 2 5(cid:13) 1 10(cid:13) 7(cid:13) (cid:13)‹(cid:13) 2.8 (cid:13)M(cid:13)o inspiraling binaries(cid:13) at Galactic center (cid:13) 5(cid:13) 3(cid:13) 10(cid:13) 20(cid:13) 30(cid:13) 40(cid:13) 60(cid:13) 50(cid:13) Distance [kpc](cid:13) 70(cid:13) 80(cid:13) 90(cid:13) 100(cid:13) FIG. 4: Contour plot of expected SNR for GW from inspiral- ing compact binaries with equal mass. Optimally-polarized GWs from an optimal direction for the detector are assumed. TAMA has the sensitivity to detect 1.4 M⊙ -1.4 M⊙ binary coalescence at Galactic center with SNR of 30. level averaged for 1 minute was kept within a few dB, typically. During the 2-week data-taking run, the noise floor-level drift was kept within 3 dB for about 90% of the total operation time. In addition, in typical cases, the noise level was easily recovered without any manual adjustment after the unlock and relock of the interfer- ometer with these automatic control systems. The noise level of the interferometer was calibrated continuously using a sinusoidal calibration signal at 625 Hz; from the amplitude and phase of this peak signal, we estimated the optical gain and cut-off frequency of the cavity. The Gaussianity of the noise level was evaluated every 30 sec- onds. We observed about 10 non-Gaussian (confidence level of 99%) events per hour in this observation run. The non-Gaussian noise will be partly removed by veto anal- ysis using other channels, such as seismic motion, laser intensity noise, contrast fluctuation, and a δL+ signal. In order to check the GW-detection ability of TAMA, 4 we calculated the expected SNR for GW from inspi- raling binaries with the interferometer noise spectrum and calculated chirp signals (Fig. 4) [19]. Here, we as- sumed optimally-polarized GWs from an optimal direc- tion for the detector. TAMA would detect GW events at the Galactic center with sufficient SNR; the SNR is about 30 in the case of chirp signals from a coales- cence of 1.4 M⊙ -1.4 M⊙ binary neutron stars at 10 kpc distance. With the burst signals from supernova explo- sions, TAMA would detect GWs with a strain amplitude of hrms ∼ 1 × 10−18 (which corresponds to a mass en- ergy of ∼ 0.01M⊙ , again at the distance to the Galactic center) with a SNR of about 10 at the frequency band from 700 Hz to 1 kHz. However, as well as a veto analysis with the other recorded channels, a coincidence analysis with other GW detectors or other astronomical channels will be required for the detection to reject non-Gaussian noise background. Conclusion. — The TAMA300 interferometer has been operated stably for over 10 hours without loss of lock, with a noise-equivalent sensitivity of h ∼ 5 × 10−21 /√Hz at the floor level. With this sensitivity and stability, TAMA has the ability to detect GW events within our galaxy, though such events are expected to be very rare. In order to increase the detection probability for GW events farther away from our galaxy, we are improving the detector sensitivity and stability further. Almost all of the noise sources which limit the detector sensitivity have been identified. The noise level will be improved with new alignment control filters and a reflective mode- matching telescope. The stability of the operation will be improved further with installation of an active isolation system and replacement of a suspension system by one with effective damping. The achieved performance of the TAMA detector is a significant milestone in the quest for direct detection of GW, and for the establishment of GW astronomy with interferometric detectors. The TAMA pro ject is supported by a Grant-in-Aid for Creative Basic Research from the Ministry of Education. ∗ e-mail: [email protected] † deceased [1] K. S. Thorne, in Three hundred years of gravitation, edited by S. Hawking and W. Israel (Cambridge Univer- sity Press, 1987), p. 330-458. [2] J. Weber, Phys. Rev. Lett. 22, 1320 (1969), Z. A. Allen et al., Phys. Rev. Lett. 85, 5046 (2000). [3] G. E. Moss, L. R. Miller, and R. L. Forward, Appl. Opt. 10, 2495 (1971). [4] A. Abramovici et al., Phys. Lett. A 218, 157 (1996), D. Shoemaker et al., Phys. Rev. D 38, 423 (1988), S. Sato et al., Appl. Opt. 39, 4616 (2000), D. I. Robertson et al., Rev. Sci. Instrum. 66, 4447 (1995), R. Takahashi et al., Phys. Lett. A 187, 157 (1994), P. Fritschel et al., Phys. Rev. Lett. 80, 3181 (1998), K. Kawabe et al., Appl. Phys. B 62, 135 (1996), M. Ando et al., Phys. Lett. A 248, 145 (1998). [5] A. Abramovici et al., Science 256, 325 (1992). [6] The VIRGO collaboration, VIRGO Final Design Report, VIR-TRE-1000-13, 1997. [7] K. Danzmann et al., Proposal for a 600m Laser- Interferometric Gravitational Wave Antenna, Max- Planck-Institut fur Quantenoptik Report 190, 1994. [8] K. Tsubono, in Gravitational Wave Experiments, edited by E. Coccia, G. Pizzella, and F. Ronga, (World Scien- tific, 1995), p. 112-114, K. Kuroda et al., in Gravitational Waves: Sources and Detectors, Edited by I. Ciufolini and F. FidecaroC (World Scientific, 1997), p. 100-107. [9] T. Nakamura, M. Sasaki, T. Tanaka, and K. S. Thorne, Astrophys. J. 487 L139 (1997). [10] S. T. Yang et al., Opt. Lett. 21, 1676 (1996). 5 [11] A. Telada, The Graduate University for Advanced Stud- ies Ph. D thesis, 1997. [12] S. Sato et al., Appl. Opt. 38, 2880 (1999). [13] R. Takahashi et al., in Gravitational Wave detection, edited by K. Tsubono, M.-K. Fujimoto, and K. Kuroda (Universal Academy Press, 1997), p. 95-102. [14] A. Araya et al., in Gravitational Wave detection, edited by K. Tsubono, M.-K. Fujimoto, and K. Kuroda (Univer- sal Academy Press, 1997), p. 55-62. [15] Y. Saito, et al., Vacuum 53, 353 (1999). [16] M. W. Regehr, F. J. Raab, and S. E. Whitcomb, Opt. Lett. 20, 1507 (1995). [17] E. Morrison, et al., Appl. Opt. 33, 5037 (1994). [18] S. Kawamura and M. Zucker, Appl. Opt. 33, 3912 (1994). [19] H. Tagoshi, et al, Phys. Rev. D (printing).
astro-ph/9504036
1
9504
1995-04-11T20:31:37
THE ORIGIN OF PLUTO'S ORBIT: IMPLICATIONS FOR THE SOLAR SYSTEM BEYOND NEPTUNE
[ "astro-ph" ]
The origin of the highly eccentric, inclined, and resonance-locked orbit of Pluto has long been a puzzle. A possible explanation has been proposed recently [Malhotra, R., {\it Nature} 365:819-21 (1993)] which suggests that these extraordinary orbital properties may be a natural consequence of the formation and early dynamical evolution of the outer Solar system. A resonance capture mechanism is possible during the clearing of the residual planetesimal debris and the formation of the Oort Cloud of comets by planetesimal mass loss from the vicinity of the giant planets. If this mechanism were in operation during the early history of the planetary system, the entire region between the orbit of Neptune and approximately 50 AU would have been swept by first order mean motion resonances. Thus, resonance capture could occur not only for Pluto, but quite generally for other trans-Neptunian small bodies. Some consequences of this evolution for the present-day dynamical structure of the trans-Neptunian region are: (i) most of the objects in the region beyond Neptune and up to $\sim\!50$ AU exist in very narrow zones located at orbital resonances with Neptune (particularly the 3:2 and the 2:1 resonances), and (ii) these resonant objects would have significantly large eccentricities. The distribution of objects in the Kuiper Belt as predicted by this theory is presented here. Keywords{solar system: origins, dynamics --- Pluto --- comets --- Kuiper Belt}
astro-ph
astro-ph
Luu, J.X. and D. Jewitt  , AJ, :- Whipple, F.L.  , Proc. Nat. Acad. Sci. :- Lyttleton, R.A.  , MNRAS, :- S. Wiggins  , Chaotic Transport in Dynamical Malhotra, R.  a, Nature, : - (Paper I) Systems, Springer-Verlag, New York Malhotra, R.  a, Physica D, : -. (Special Williams, J.G. and G.S. Benson  , AJ, :- issue on \Modeling the Forces of nature") Wisdom, J. and M.J. Holman  , AJ, :- Malhotra, R.  b, Cel. Mech.&Dyn. Ast., :-   Malhotra, R.  , LPSC, XXVI:- Malhotra, R. and J.G. Williams  , in Pluto & Charon, D. Tholen & S.A. Stern, eds., Univ. of Arizona Press, Tucson. (in press) B.G. Marsden  a, IAU Circular   B.G. Marsden  b, IAU Circular  W.B. McKinnon and S. Mueller  , Nature, :- McKinnon, W.B. and E.M. Parmentier  , in Satel- lites, J.A. Burns & M.S. Matthews, eds. Univ. of Arizona Press, Tucson. p. { Milani, A., A.M. Nobili and M. Carpino   , Icarus, :- Nobili, A.M., A. Milani and M. Carpino   , A&A, :- Olsson-Steel, D.I.  , A&A,  :- S.J. Peale  , in Satel lites, J. Burns & M. Matthews, eds. Univ. of Arizona Press, Tucson. p.  - Quinn, T.R., S. Tremaine and M.J. Duncan  , ApJ, :- Standish, E. M.  , AJ, :- Sussman, G.J. and J. Wisdom  , Science, :-  G. Tancredi and J.A. Fernandez  , Icarus, : -  Tombaugh, C.W.  , in Planets and Satel lites, G.P. Kuiper & B.M. Middlehurst, eds. Univ. Chicago Press, Chicago. pp. - Weissman, P.R.  , Nature, :- macros v.. This -column preprint was prepared with the AAS L T X A E 
astro-ph/0412063
1
0412
2004-12-02T17:16:47
Bending Instability of Stellar Disks: The Stabilizing Effect of a Compact Bulge
[ "astro-ph" ]
The saturation conditions for bending modes in inhomogeneous thin stellar disks that follow from an analysis of the dispersion relation are compared with those derived from $N$-body simulations. In the central regions of inhomogeneous disks, the reserve of disk strength against the growth of bending instability is smaller than that for a homogeneous layer. The spheroidal component (a dark halo, a bulge) is shown to have a stabilizing effect. The latter turns out to depend not only on the total mass of the spherical component, but also on the degree of mass concentration toward the center. We conclude that the presence of a compact (not necessarily massive) bulge in spiral galaxies may prove to be enough to suppress the bending perturbations that increase the disk thickness. This conclusion is corroborated by our $N$-body simulations in which we simulated the evolution of almost equilibrium, but unstable finite-thickness disks in the presence of spheroidal components. The final disk thickness at the same total mass of the spherical component (dark halo + bulge) has been found to be much smaller than that in the simulations where a concentrated bulge is present.
astro-ph
astro-ph
Journal-Ref: Astronomy Letters, 2005, v. 31, No 1, pp. 15-29 Bending Instability of Stellar Disks: The Stabilizing Effect of a Compact Bulge N. Ya. Sotnikova and S. A. Rodionov Astronomical Institute, St. Petersburg State University, Universitetskii pr. 28, 198904 Russia Received January 19, 2004 Abstract The saturation conditions for bending modes in inhomogeneous thin stellar disks that follow from an analysis of the dispersion relation are compared with those derived from N-body simulations. In the central regions of inhomogeneous disks, the reserve of disk strength against the growth of bending instability is smaller than that for a homogeneous layer. The spheroidal component (a dark halo, a bulge) is shown to have a stabilizing effect. The latter turns out to depend not only on the total mass of the spherical compo- nent, but also on the degree of mass concentration toward the center. We conclude that the presence of a compact (not necessarily massive) bulge in spiral galaxies may prove to be enough to suppress the bending perturbations that increase the disk thickness. This conclusion is corroborated by our N-body simulations in which we simulated the evolu- tion of almost equilibrium, but unstable finite-thickness disks in the presence of spheroidal components. The final disk thickness at the same total mass of the spherical component (dark halo + bulge) has been found to be much smaller than that in the simulations where a concentrated bulge is present. Key words: galaxies, stellar disks, bending instabilities. 4 0 0 2 c e D 2 1 v 3 6 0 2 1 4 0 / h p - o r t s a : v i X r a 1 1 INTRODUCTION A peculiarity of the disks in spiral galaxies is that these are rather thin objects. Their thickness is several times smaller than the radial scale length. How far stars can go from the principal galactic plane due to their vertical random velocity component determines the disk thickness at fixed star surface density. The larger the velocity dispersion, the thicker the disk. The random velocities of young stars are known to be low, but the stellar ensemble can subsequently heat up through various relaxation processes; i.e., the random velocity dispersion can increase. Thus, the stellar disk thickness depends on how effective the relaxation processes are in galaxies, and it is ultimately determined by the factors that suppress or trigger the various heating mechanisms. Three basic stellar disk heating mechanisms are commonly discussed in the literature: the scattering of stars by giant molecular clouds (Spitzer and Schwarzschild 1951, 1953), the scattering by transient spiral density waves (among the first results of numerical simulations are those obtained by Sellwood and Carlberg (1984)), and the heating of the ensemble of stars that constitute the disks of spiral galaxies as they interact with external sources, for example, with low- mass satellites (see, e.g.,Walker et al. 1999; Velasquez and White 1999). A second remarkable peculiarity of the stellar disks is that their structure is unusually "fragile". This peculiarity was revealed by a linear analysis of the collisionless Boltzmann equation and has been repeatedly illustrated by N-body simulations. Numerous studies have shown that the initially regular structure of the stellar disks can change radically due to the growth of various instabilities, which give rise to large-scale structures both in the disk plane (bars, spiral arms, rings) and in the vertical direction (warps). The analytically and numerically obtained saturation conditions for unstable modes impose the most severe restrictions on the global structural and dynamical parameters of the stellar disks. A local analysis suggests that the stars at a given distance R from the disk center must have a radial velocity dispersion σR(R) larger than some minimum critical value σcr R (R) (Toomre 1964) for the disk to be gravitationally stable in the region under consideration (at least against the growth of axisymmetric perturbations). In addition, for the disk to be stable against the growth of bending perturbations, the ratio of the vertical and radial velocity dispersions σz/σR must also be larger than some value approximately equal to 0.2 − 0.37 (Toomre 1966; Kulsrud et al. 1971; Polyachenko and Shukhman 1977; Araki 1985). The latter quantity determines the minimum thickness of the galactic disk, and its comparison with the observed value allows the contribution of the bending instability to the vertical disk heating to be estimated and compared with the contribution of other relaxation mechanisms. On the other hand, if we exclude the heating mechanisms mentioned above from our analysis and take, as is commonly done, the condition for marginal disk stability against the growth of bending modes by fixing σz/σR at a level of the linear approximation, 0.29 − 0.37, then the velocity dispersion σz at the same star surface density will decrease with decreasing σR (see, e.g., Zasov et al. 1991). In this case, the disk will have a smaller thickness. The presence of a spheroidal component, for example, a dark halo is known to produce a stabilizing effect and to decrease the minimum value of σcr R required for gravitational stability. Consequently, the disks embedded in a massive halo, on average, must have low 2 values of σz and be, on average, thinner. Based on similar reasoning, Zasov et al. (1991, 2002) showed that the relative disk thickness z0/h (z0 is the half-thickness of the disk, and h is the radial exponential scale length) is proportional to Md/Mt, where Md and Mt are, respectively, the mass of the disk and the total mass of the galaxy within a fixed radius. These authors also concluded that a small disk thickness suggests the existence of a massive dark halo in the galaxy. Moreover, based on N-body simulations, Zasov et al. (1991) and Mikhailova et al. (2001) constructed a dependence that allows the relative mass of the dark halo Mh/Md to be estimated from z0/h. However, as we show below, the relationship between the relative disk thickness and the mass of the spheroidal component is more complex. In this paper, we numerically analyze the saturation conditions for bending instability in inhomogeneous three-dimensional stellar disks at nonlinear stages in the presence of a spheroidal component of different nature (a stellar bulge and a dark halo) and the constraints imposed on the final disk thickness. 2 PECULIARITIES OF THE GROWTH OF BENDING INSTABILITY IN INHOMOGE- NEOUS THIN STELLAR DISKS 2.1 Global Modes To understand how the growth of bending instability in inhomogeneous disks differs from that in homogeneous disks, let us first turn to the result of Toomre (1966). Toomre was the first to derive the dispersion relation for long-wavelength bending perturbations in an infinitely thin gravitating layer with a nonzero velocity dispersion ω2 = 2πGΣk − σ2 xk2 . (1) where Σ is the star surface density of the layer, and σ2 x is the velocity dispersion along a particular coordinate axis in the plane of the layer. It follows from Eq. (1) that the perturbations with a wavelength λ = 2π/k > λJ = σ2 x/GΣ are stable, since ω2 > 0 in this range. Relation (1) was derived for an infinitely thin disk and, when applied to finite-thickness disks, is valid only for perturbations with a wavelength longer than the vertical scale height of the system, i.e., for λ >> z0, where z0 is the half-thickness of the layer. It can be shown from general considerations that the longest-wavelength perturbations in finite-thickness disks with a wavelength λ < λ2 ≈ z0 are stable, since they are smeared by thermal motions in the plane of the layer (see, e.g., Polyachenko and Shukhman 1997). Having derived the corresponding dispersion relation, Polyachenko and Shukhman (1977) were the first to find the exact location of the stability boundary in the short-wavelength range for a homogeneous flat finite-thickness layer. Araki (1985) (see also Merritt and Sellwood 1994) obtained a similar result for a homogeneous layer with a vertical density pro?le close to the observed one in real galaxies1 ρ(R, z) = ρ(R, 0)sech2(z/z0). As regards the 1This profile corresponds to the model of an isothermal layer (Spitzer 1942) and describes well the σx σz vertical density variations observed in galaxies (van der Kruit and Searle 1981). 3 intermediate-wavelength (λ2 < λ < λJ) perturbations, they are unstable, as follows from the results of the studies mentioned above. As the disk thickness z0 increases, the wavelength λ2 increases and tends to λJ. When λ2 = λJ, the disk is stabilized against bending perturbations of any wavelengths. The following analytical estimate in the linear approximation obtained both from qualitative considerations (Toomre 1966; Kulsrud et al. 1971) and from an accurate analysis of the dispersion relation for a finite-thickness layer (Polyachenko and Shukhman 1977; Araki 1985) is valid: (σz/σx)cr ≈ 0.29 − 0.37 . (2) The instability is completely suppressed if σz/σx > (σz/σx)cr and grows if σz/σx < (σz/σx)cr. As regards the inhomogeneous models, the first thing that radically distinguishes the growth of bending instability in inhomogeneous disks from that in a homogeneous layer is the existence of global unstable bending modes with a wavelength longer than the disk scale length. This conclusion follows from an analysis of the equation that describes the evolution of long-wavelength bending perturbations in an in infinitely thin disk with a ra- dially decreasing density (Polyachenko and Shukhman 1977; Merritt and Sellwood 1994). For special disk models, it was shown that the region of stable long-wavelength pertur- bations narrows significantly in this case (see, e.g., Fig. 2 from Merritt and Sellwood (1994)). This is attributed to the fact that the restoring force from the perturbation that grows in an inhomogeneous disk (Merritt and Sellwood 1994) or in a radially bounded disk (Polyachenko and Shukhman 1977) is always weaker than the corresponding restor- ing force in a homogeneous infinite layer. This fact was demonstrated more clearly by Sellwood (1996), who noted that the dispersion relation (1) could be used to analyze the bending instability in inhomogeneous disks (at least qualitatively) if another term related to the restoring force from the unperturbed disk is added to it: ω2 = ν2 d + 2πGΣk − σ2 xk2 , (3) where νd = p∂2Φd(R, z)/∂z2 is the vertical oscillation frequency of the stars, and Φd(R, z) is the potential of the disk. For disks with a nonflat rotation curve, the additional term ν2 d can play a destabilizing role. As was noted by Sellwood (1996), for an infinitely thin inhomogeneous disk, ν2 d = ∂2Φd(R, z) ∂z2 = − 1 R ∂ ∂R (cid:18)R (cid:12)(cid:12)(cid:12)(cid:12)z=0+ 1 R dv2 c,d dR , = − (4) ∂2Φd ∂R (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)z=0 (vc,d is the circular rotational velocity of the disk), and ν2 d < 0 for dv2 c,d/dR > 0. Thus, an additional expulsive force from the unperturbed disk emerges in the central regions where the rotation curve rises. The destabilizing effect of the disk in the central regions gives rise to another region at small wave numbers (long λ) where ω2 < 0. This region is responsible for the growth of largescale bending instability and the emergence of global modes. In this case, one might expect a larger disk thickness and a larger value σz/σx (σz/σR) than those in homogeneous models to be required to suppress the instability. Applying all of this reasoning to finite-thickness disks is not quite obvious. On the one hand, it may be assumed that the restoring force from the unperturbed disk Fz = 4 −∂Φd/∂z for R → 0 behaves as Fz ≃ −GMdz/z3 (here, Md is the total mass of the disk) starting from some z, i.e., decreases (in magnitude) with increasing z. Consequently, ν2 d = ∂2Φd/∂z2 = −∂Fz/∂z becomes negative (Sellwood 1996). the general formula On the other hand, we can exactly calculate the z dependence of ν2 d for a given R from ν2 d(r) = −∂Fz/∂z = − ∂ ∂z Z Gρd(r′)(z′ − z) r′ − r3 d3r′ , for the density profile that is commonly used to describe the disks of spiral galaxies: ρd(R, z) = Md 4πh2z0 · exp(cid:18)− R h(cid:19) · sech2(cid:18) z z0(cid:19) (5) (6) where h is the exponential scale length of the disk, z0 is the vertical scale height, and Md is the total mass of the disk. The derived dependence2 for various R is shown in Fig. 1b. Figure 1a shows the z dependence of the vertical force Fz, which necessarily has d is negative at z >∼2z0 (on the periphery, an extremum at some z. In the central regions, ν2 the passage to the negative region occurs at even larger z/z0). This implies that all of the above reasoning for finitethickness disks is directly applicable only to largeamplitude perturbations. It is also clear that the bellshaped (or axisymmetric) mode with the azimuthal number m = 0 must be most unstable in the central regions. If the amplitude of the axisymmetric bend increases significantly during the growth of bending instability, then it can raise the central stars above the plane of the disk and bring them into a "dangerous" zone where ν2 d < 0; subsequently, the self-gravity of the disk itself will contribute to the growth of a bend to the point of saturation. 2.2 The Central Regions of Hot Infinitely Thin Disks The second peculiarity of the growth of bending instability in inhomogeneous infinitely thin disks is as follows. The degree of disk instability against bending perturbations depends on the degree of disk "heating" in the plane. Exact and approximate analyses of the dispersion relation for an inhomogeneous infinitely thin disk (Polyachenko and Shukhman 1977; Merritt and Sellwood 1994) show that the modes with increasingly large azimuthal numbers in the central regions become unstable for all wavelengths as the fraction of the kinetic energy contained in the random motions in the disk plane increases. The conclusion about the existence of a Toomre parameter QT (Toomre 1964), which characterizes the degree of heating of the stellar disk in the plane, at which the disk cannot be stabilized against bending perturbations of any wavelengths can also be drawn from Eq. (3), i.e., the equation written for long-wavelength perturbations in a homogeneous layer, but "rectifided" by the term ν2 d to include the inhomogeneity effects. 2We see from Eq. (5) that triple integrals over infinite intervals must generally be calculated to determine ν 2 In our integration,we used the gsl library (information about the gsl (GNU Scientific Library) project can be found at http://sourses/redhat/com/gsl). For models with known analytical density-potential pairs, for example, the Miyamoto-Nagai disk (see Binney and Tremaine (1987), p. 44), a comparison of the analytical z dependence of ν 2 d at a given R with that calculated numerically by triple integration yielded a close match. d. This can be easily done by using adaptive algorithms for calculating the integrals. 5 Indeed, ω2 < 0 for any k if ν2 d + (πGΣ)2 σ2 x < 0 . Given (4), where Ωd = vc,d R is the angular velocity, and κ2 ν2 d = 2Ω2 d − κ2 d , d = 2 the epicyclic frequency. Then, ω2 < 0 2Ω2 d − κ2 d + (πGΣ)2 σ2 x < 0 . v2 c,d R2 (cid:18)1 + R vc,d dvc,d dR (cid:19) is the square of (7) (8) For the central regions of a rigidly rotating disk (κ2 d = 4Ω2 ω2 < 0 σx > √2 πGΣ κd ≈ d), we obtain from (7) √2 σcr R , or ω2 < 0 QT > √2 for any wavelength, where QT = σx/σcr R is the Toomre parameter3 (Toomre 1964). Condition (8) is not an exact criterion; it is only an estimation relation. However, it shows that the central regions of hot stellar disks (QT >> 1) with a large reserve of strength against the growth of instabilities in the plane of the disk (bars, spiral arms) cannot be stabilized against the growth of bending perturbations of any wavelengths. The theory constructed for a homogeneous infinitely thin layer does not yield this regime. It should be borne in mind, however, that the contribution of the destabilizing term (ν2 d) must decrease with increasing disk thickness; therefore, the instability will be saturated at a large, but finite disk thickness. However, the following might be expected: other things being equal, the hotter the initial model in the plane, i.e., the larger the Toomre parameter QT, the higher the saturation level. 3 BENDING INSTABILITY: NUMERICAL SIMU- LATIONS OF THREE-DIMENSIONAL DISKS The nonlinear growth stages of bending instability in inhomogeneous finite-thickness stel- lar disks have been extensively investigated by numerically solving the gravitational N- body problem for various stellar disk models. Raha et al. (1991) first observed the bending instability of bars in their numerical simulations; Sellwood and Merritt (1996) and Merritt and Sellwood (1994) studied the nonlinear regime of bending instability in nonrotating disks with the radial density profiles that corresponded to the Kuzmin -- Toomre model (see, e.g., Binney and Tremaine (1987), p. 43); Griv et al. (1998) numerically analyzed the development of a bend in a layer of newly formed stars; Tseng (2000) simulated the evolution of the vertical structure of a homogeneous, initially thin disk of finite radius; 3Actually, the Toomre parameter should have been defined as σR/σcr R , but the difference does not matter in our case. 6 Sotnikova and Rodionov (2003) considered a rotating disk with an exponential density profile along the R axis and assumed the presence of a dark halo in the system. The latter authors analyzed the question of how the evolution of initially equilibrium thin disks depends on the governing parameters of the bending instability, which include the initial disk half-thickness z0, the Toomre parameter QT, and the relative mass of the dark halo Mh/Md within a fixed radius. Below, we list the most important conclusions that follow from the N-body simulations described in the literature. These conclusions in many respects agree with those given in the section entitled "Peculiarities of the growth of bending instability. . . " of this paper for inhomogeneous infinitely thin disks. (1) In inhomogeneous models, all of the experimentally observed modes are global, i.e., the scale length of the unstable perturbations is larger than the disk scale. The linear theory constructed for homogeneous models yields no such result. (2) The saturation level for the bending instability depends on σR (or QT). The larger the reserve of disk strength against perturbations in the disk plane, the greater the dificulty to stabilize the disk against the growth of bending perturbations. A rapid and significant (occasionally catastrophic) increase in the disk thickness, particularly in the central regions, to values that are severalfold larger than those yielded by a linear analysis for homogeneous, moderately hot models (Toomre 1966; Kulsrud et al. 1971; Polyachenko and Shukhman 1977; Araki 1985) was observed in all of the simulations with initially hot disks (Sellwood and Merritt 1994; Tseng 2000; Sotnikova and Rodionov 2003). This mechanism may be responsible for the formation of central bulges in spiral galaxies, at least it can feed the spherical component with new objects. (3) The central regions of the disk are most unstable (Sellwood and Merritt 1994; Merritt and Sellwood 1994; Griv and Chiueh 1998; Griv et al. 2002; Sotnikova and Rodi- onov 2003). It is here that the bending modes are formed. Subsequently, their amplitude increases, sometimes significantly. This is particularly true for the perturbations with m = 0. The nonaxisymmetric modes (with the azimuthal numbers m = 1 and m = 2) drift to the disk periphery, temporarily creating the effect of a largescale warp of the entire galaxy, and are then damped4. The instability in the central regions of the stellar disks is saturated at (Sotnikova and Rodionov 2003) σz/σR ≈ 0.75 − 0.8 . (4) The presence of a massive dark halo, which was included in the models by Sotnikova and Rodionov (2003), has always been a stabilizing factor that suppresses the growth of bending modes. This effect appears to have been first described qualitatively by Zasov et al. (1991). 4In contrast to the result of Griv et al. (2002), the largescale S-shaped or U-shaped warp of the galactic edge always disappeared on long time scales (¿ 5 Gyr) in all of the simulations by Sotnikova and Rodionov (2003). 7 4 THE STABILIZING EFFECT OF THE SPHEROIDAL COMPONENT: A QUALITA- TIVE ANALYSIS The remarkable agreement (at least on a qualitative level) between the conclusions that follow from the analysis of the dispersion relation for an infinitely thin disk and the results of numerical simulations with three-dimensional disks allows us to analyze the applicability of yet another conclusion that can be drawn from Eq. (3) to finite-thickness disks. Since the central regions of the disk are most unstable, it is important to separate out the factors that have a stabilizing effect precisely on these regions. An additional spherical component (a dark halo, a bulge) can be such a stabilizing factor. In this case, another term related to the restoring force exerted from the spherical component appears in Eq. (3), with ν2 sph = ∂2Φsph(r) ∂z2 (cid:12)(cid:12)(cid:12)(cid:12)z=0 > 0 . (9) This term was also introduced by Sellwood (1996) when analyzing the dispersion relation for the longwavelength bending perturbations of an infinitely thin disk, but its role was not studied. Zasov et al. (1991, 2002) and Mikhailova et al. (2001) concluded that the minimum possible relative thickness of an equilibrium stellar disk, z0/h, decreases with increasing relative mass of the dark halo. Let us consider the relationship between the stellar disk thickness and the mass of the spheroidal component in terms of the stabilization conditions for the bending modes in inhomogeneous thin disks. We take specific bulge and halo models (these models were subsequently used in our numerical calculations). A Plummer sphere is taken as the bulge model. Its potential is (see, e.g., Binney and Tremaine (1987), pp. 42 -- 43) Φb(r) = − GMb (r2 + a2 b)1/2 , (10) where Mb is the total mass of the bulge, and ab is the scale length of the matter distri- bution. We describe the potential of the dark halo in terms of the logarithmic potential (see, e.g., Binney and Tremaine (1987), p. 46) Φh(r) = ∞ v2 2 ln(r2 + a2 h) + const , (11) where ah is the scale length, and v∞ is the velocity of a particle in a circular orbit of infinite radius. The parameter v∞ is related to the mass of the halo with a sphere of given radius r by Mh(r) = r3 ∞ v2 G ). r2 + a2 h For the additional stabilizing term in the dispersion relation, the models of the spher- ical components (the bulge and the halo) that we used yield ν2 b = GMb (R2 + a2 b)3/2 ; ν2 h = v2 ∞ R2 + a2 h . 8 The stabilizing effect of the spherical component weakens at large R, but the disk itself in the peripheral regions has a stabilizing effect: Fz ≈ −GMdz/R3 at z << R ⇒ ν2 d = −∂Fz/∂z > 0. On the other hand, the strength of the effect increases in the central (most unstable) regions (i.e., for R → 0); this strength depends not only on the total mass of the spherical component (Mb or v∞), but also on the degree of matter concentration toward the center, ab and ah. It thus follows that the presence of a compact (not necessarily massive) bulge in galaxies may prove to be enough to suppress the bending perturbations. This implies that the disks of galaxies with compact bulges can be as thin as the disks embedded in a massive dark halo. To test our conclusion, we carried out a series of numerical simulations. 5 THE STABILIZATION OF BENDING PERTUR- BATIONS BY A COMPACT BULGE: N -BODY SIMULATIONS 5.1 The Method We used an algorithm based on the hierarchical tree construction method (Barnes and Hut 1986) to simulate the evolution of a self-gravitating stellar disk. In our calculations, we always included the quadrupole term in the Laplace expansion of the potential produced by groups of distant bodies. The parameter θ (Hernquist 1993) that is responsible for the accuracy of calculating the gravitational force was chosen to be 0.7 in all our simulations. The NEMO software package (http://astro.udm.edu/nemo; Teuben 1995) was taken as the basis. We enhanced the capabilities of this package by including several original codes for specifying the equilibrium initial conditions in a flat stellar system5 and supplemented it with new codes that allow us to easily analyze the data obtained and to present them in convenient graphical and video formats. 5.2 The Numerical Model When constructing the galaxy model, we distinguished two components in it: a self- gravitating stellar disk and a spherically symmetric component that was described in terms of the external static potential, which is a superposition of two potentials, (10) for the bulge and (11) for the dark halo. At large distances R from the center of the stellar system, in the region where the halo dominates, potential (11) yields a flat rotation curve. The disk was represented by a system of N gravitating bodies with the density profile (6). The initial conditions in the N-body problem suggest specifying the mass, position in space, and three velocity components for each particle. The particle coordinates are naturally determined in accordance with the disk matter density profile (6); the distant regions of the disk are disregarded. We took only those particles for which the cylindrical radius R < Rmax and z < zmax. The mass of all particles was assumed to be the 5The mkexphot code for specifying equilibrium stellar disk models from the NEMO package has limitations on the parameters of the outer halo and does not enable the gravitational field of the bulge to be specified. The models constructed using our codes closely agree with those obtained using the mkexphot code for identical parameters. 9 same. The total mass of the particles was equal to the mass of the disk region under consideration (i.e., the disk region for which R < Rmax and z < zmax). The particle velocities were specified using the equilibrium Jeans equations by the standard technique (see, e.g., Hernquist (1993), Section 2.2.3). 5.3 Specifying the Velocity Field in the Model Galaxy To specify the initial particle velocities for a disk that is in equilibrium in the plane and in the vertical direction, we make the following assumptions: (1) The velocity distribution function is the Schwarzschild one; in other words, the particle velocity distribution function has only four nonzero moments: the mean azimuthal velocity ¯vϕ, the radial velocity dispersion σR, the azimuthal velocity dispersion σϕ, and the vertical velocity dispersion6 σz. (2) All four moments depend only on the cylindrical radius R and do not depend on z. (3) The epicyclic approximation is valid7. (4) σ2 R is proportional to the surface density of the stellar disk, i.e., σR ∝ exp (−R/2h) (this assumption agrees well with the observational data; see, e.g., van der Kruit and Searle 1981). The following relations for the moments (in which the R dependence was omitted for simplicity) can then be derived from the Jeans equations (see, e.g., Binney and Tremaine 1987): ϕ + c + σ2 ϕ = v2 ¯v2 R − σ2 κ , σϕ = σR 2Ω σ2 z = πΣdz0 , R Σd ∂Σdσ2 R ∂R , (12)   where vc is the circular velocity of a particle placed in the total potential of the disk and ∂Φb the spherical component, the bulge and the halo, (v2 ∂R c,d + v2 c,b = R c,h; v2 ; c = v2 is the angular velocity, and κ = r2 c,b + v2 dv2 c dR 1 R v2 c R2 + is the epicyclic ∂Φh ∂R v2 c,h = R frequency. ), Ω = vc R The circular velocities for the bulge (10) and the halo (11) have analytical expressions. The circular velocity for the disk (6) can be determined by numerical integration (see the section entitled "Global modes") using the general formula v2 c,d(r) = Z Gρd(r′) · (r′ − r) · R r′ − r3 d3r′ , (13) where R is the projection of the vector r onto the disk plane. The Jeans equations that are used to derive relations (12) are known to provide no exact disk equilibrium (see, e.g., Binney and Tremaine 1987). Moreover, the last relation in (12), which follows from the vertical equilibrium condition for a disk with the density 6As many members of the astronomical society, we have the bad habit of calling the standard of the distribution function the dispersion. 7In the central regions of the disk, this approximation breaksdown. 10 profile (6) and a z-independent σz, was written without including the influence of the additional spheroidal components. The adjustment to equilibrium occurs on time scales of the order of several vertical oscillation times 1/νd. This time was always no longer than 100 − 120 integration time steps for the equations of motion (the thinner the disk, the shorter this time) and much shorter than the instability growth time scale in the disk. Moreover, in the context of the problem of the growth and saturation of unstable modes, a small deviation of the disk from equilibrium at the initial time may be treated as an additional initial perturbation. The fourth assumption (see above) about the radial velocity dispersion σ2 R(R) causes dificulties in calculating ¯vϕ in the central regions. In the first equation of system 12, ¯v2 ϕ is occasionally negative at small R (since σ2 R(R) rapidly increases toward the center, the last term on the right-hand side can make a large negative contribution). For this reason, the dependence for σR was reduced at the center (Hernquist 1993): σR ∝ exp(cid:16)−pR2 − 2a2 s/2h(cid:17) . (14) If the parameter as is taken to be h/4− h/2, then this proves to be enough to properly calculate ¯vϕ. In the central regions of the disk, σz adjusted to σR in such a way that the ratio σz/σR was constant at a given half-thickness z0 = const in the initial model throughout the disk. The proportionality factor in (14) can be determined via the Toomre parameter QT at some radius Rref: σR(Rref) = QT σcr R (Rref) = QT 3.36 · Σd(Rref) κ(Rref ) . (15) The sought proportionality factor can be obtained from (14) and (15). Specifying QT, which ensures disk stability in the plane, at Rref ≈ 2.5h yields the condition QT(R) ≥ QT(Rref) (Hernquist 1993). The latter, in turn, ensures a stability level against perturbations in the disk plane no lower than that at Rref. The initial half-thickness z0 for the adopted QT(Rref) was chosen in such a way that ?z/?R was less than 0.3 − 0.4, which ensured initial instability against the growth of bending modes. 5.4 Parameters of the Problem All of the results discussed below are presented in the following system of units: the gravitational constant is G = 1, the unit of length is Ru = 1 kpc, and the unit of time is tu = 1 Myr. The unit of mass is then Mu = R3 u = 22.2 · 1010M⊙, and the unit of velocity is vu = Ru/tu = 978 km s−1. The number of bodies in the simulations was N = 300 000 (in several cases, 500 000 and 600 000). The force of interaction between two particles with coordinates ri and rj and masses mi and mj was modified, as is commonly done, as follows: u/Gt2 Fij = Gmimj rj − ri (rj − ri2 + ǫ2)3/2 , where ǫ is the softening length of the potential produced by an individual particle. When collisionless systems are simulated, this parameter is introduced for two reasons. First, the 11 divergence of the interaction force in close particle -- particle encounters must be avoided when integrating the equations of motion. Second, when the phase density of a collisionless system is represented by a finite number of particles, the inevitable fluctuations in the particle distribution must be smoothed in such a way that the forces acting in the system being simulated were are to the forces acting between the particles in a system with a smoother density profile. The softening length ǫ was chosen to be 0.02. This value is approximately a factor of 2 or 3 smaller than the mean separation between the particles (at N = 300 000) within the region containing half of the disk mass. On the one hand, it matches the criterion for choosing ǫ based on minimization of the mean irregular force (Merritt 1996) and, on the other hand, allows the vertical structure of thin disks to be adequately resolved. To integrate the equations of motion for particles in the self-consistent potential of the disk and the external field produced by the spheroidal component, we used a leapfrog scheme that ensured the second order of accuracy in time step. The time step was 0.5 (in several models, 0.25)8. We constructed a total of about 60 models. The entire set of models can be arbitrarily divided into two classes: the models with and without bulges. The scale length of the density distribution in the bulge ab was assumed to be equal to 0.5 almost for all of the models with bulges. In several models without bulges, we chose a concentrated halo (ah = 2). In all of the remaining cases, the halo was "looser" (ah = 10). The total relative mass of the spheroidal components µ = Msph(4h)/Md(4h) was varied over the range 0.25 to 3.5 The disk in our models has the following parameters9: h = 3.5 and the disk mass (in dimensional units) Md(4h) = (4 − 8) × 1010M⊙. The initial thickness was varied over the range z0 = 0.1 − 0.5. In order not to abruptly cut off the model disk at the radius corresponding to the optical radius (∼ 4h), we chose Rmax = 25, with zmax = 5. The smoothing parameter of the initial radial profile of the velocity dispersion σR is as = 1. The parameter QT in the discussion of our simulations is given for the radius Rref = 8.5. 5.5 Simulation Results Previously (Sotnikova and Rodionov 2003), we showed that there are two distinct vertical stellar disk relaxation mechanisms related to bending instability: the bending instability of the entire disk and the bending instability of the bar forming in the disk. The for- >∼2.0), and the bar mer mechanism dominated in galaxies that are hot in the plane (QT formation was suppressed in this case; the latter mechanism dominated in galaxies with a moderate Toomre parameter QT (such galaxies were unstable against the growth of a bar mode). In the simulations whose results are presented and analyzed below, we also considered two distinct cases: hot disks (QT = 2.0) in which only bending instability developed, and cooler models (QT = 1.5) -- here, we observed the combined effect of the two types of instability. Hot disks. For hot (in the plane) disks (QT = 2.0), we revealed distinct patterns of growth and saturation of bending perturbations that are consistent with the conclusions 8The choice of the time step is limited above by ǫ -- the particle must take at least one step on the smoothing length. 9These parameters are close those of the disk in our Galaxy. 12 following from a qualitative analysis of the dispersion relation (3). The stabilizing effect from the presence of a massive spherical component that was discussed in the section entitled "The stabilizing effect of the spheroidal component. . . " is clearly seen in Fig. 2. This figure shows the radial profile of σz/σR for model 12 with Mb = 0 and µ = 3.0. Throughout the disk, σz/σR was set at ≈ 0.35 and was determined not by the bending instability, but by the disk heating through the scattering of stars by inhomogeneities related to the different disk thickness in the spiral arms and in the interarm space (Sotnikova and Rodionov 2003). We will demonstrate the stabilizing effect of a compact (not necessarily massive) bulge comparable to the effect of a massive dark halo with a broader density profile using the results obtained for the following group of models as an example: 50, 76, 75, 49, and 53. In all five models, the total mass of the spheroidal component is the same and accounts for half of the diskmass within four exponential disk scale lengths, µ = 0.5, but it is differently distributed between the two spherical subcomponents. In model 50, all of the mass is contained in the halo (µ = µh = Mh(4h)/Md(4h)); in model 53, only a compact bulge is present10 (µ = µb = Mb(4h)/Md(4h)). The remaining models are intermediate between the two extreme models. The initial thickness for all of the models was chosen to be the same, z0 = 0.1. The rotation curves for these models are shown in Fig. 3. The variety of the shown curves to some extent reflects the actual variety of rotation curves for spiral galaxies. We traced the evolution of these models up to t = 5000. Figure 4 illustrates the variations in the dynamical parameters of the disk σR and σz the radial and vertical velocity dispersions calculated at R = 2h. All of the models demonstrate an initial increase in σz and a decrease in σR. Subsequently (after t ≈ 1000), the latter parameter reaches an approximately constant value, while σz for some of the models (this is primarily true for the model with a massive bulge) continues to slowly increase. The number of particles in our models and the softening length were chosen in such a way that the two- body relaxation time was much longer than the time scale on which we considered the evolution of our numerical models. The absence of heating related to numerical relaxation is confirmed by the behavior of σR and the preservation of the pattern of evolution of the system as the number of particles increases to N = 600000. The continuing small secular increase in the vertical velocity dispersion probably reflects the fact that some of our models did not reach a steady state11. Figure 5 shows five frames that correspond to the late evolutionary stages of our model disks. As expected, the saturation level for the bending instability in model 50 was very high and did not match the standard linear criterion. The galaxy greatly thickened at the final evolutionary stages. However, when we transferred 50% of the mass from the halo to the bulge (model 49: µh = µb = 0.25), the picture changed. The saturation level for the bending instability became much lower. At the final evolutionary stages, the disk was much thinner than that in model 50. In model 53, when we placed all of the mass of the spherical component in the bulge, the amplitude of the observed bend was very low, and the galaxy remained quite thin even at the late evolutionary stages. 10In model 53, a bulge with a mass equal to half of the disk mass and a scale length of 500 pc is atypical of real galaxies. We consider this as a limiting case. 11If the system has no third integral of motion, then its evolution to equilibrium must eventually lead to the relation σz = σR. 13 The disk thickness can be quantitatively estimated as the root-mean-square (rms) value of the z coordinates of the disk particles, zrms = √< z2 > − < z >2. This estimate profile (6), the relationship between this parameter and z0 is given by zrms = π/2√3z0 ≈ is commonly encountered in the literature. It can be shown that for the vertical density 0.91z0. In practice, however, this parameter proved to be a not very good characteristic of the thickness. First, the fluctuations in this parameter along R were found to be great even when using a large number of particles if only no averaging is performed in concentric rings of large width. Second, the thickness calculated in this way turns out to be systematically overestimated due to the existing of a significant tail of the particles that went far from the disk plane. For these reasons, we estimated the disk thickness at a given distance R through the median of the absolute value of z that was designated as z1/2. Twice the value of z1/2 is nothing but the disk thickness within which half of the particles is contained. For the density profile (6), 2z1/2 = z0 · ln 3 ≈ 1.1z0. Figure 6 shows the differences between the radial disk thickness profiles for model 53 obtained by the two described methods (the averaging was performed in concentric rings; the ring width was ∆R = 0.4). We see that z1/2 behaves much more smoothly (we have in mind the overall monotonic dependence of the density and the fluctuation level) than does the rms value of the z coordinate commonly used to estimate the disk thickness in N-body simulations. Figure 7 shows the radial disk thickness profile for models 50, 76, 75, 49, and 53 at the time t = 3000. Note that the thickness profile for model 50 is rather unusual in shape. This shape is most likely attributable to the existence of X-shaped stationary orbits in the central regions that arise at a certain disk thickness when there are conditions for the resonance between the stellar oscillation frequencies in the disk plane and in the vertical direction (see, e.g., Patsis et al. 2002). We see the following from Fig. 7, as well as from Fig. 5, which show the edge-on views of the model galaxies: the thinner the galactic disk, the larger the mass of the spheroidal component contained in a compact bulge. Since not all of our models reached a steady state, their thickness continues to slowly increase (Fig. 8), but the differences in thickness are always preserved. Similar results were obtained for all of the remaining such models with the same mass of the spherical component in the range Msph(4h) = 0.25Md(4h) to Msph(4h) = 3.5Md(4h). The stabilizing effect of a bulge was particularly pronounced in those cases where the bulk of the galactic mass was contained in the disk. The final disk thickness at fixed initial QT was determined only by the relative mass of the spheroidal component and the contribution of the bulge to this mass and did not depend on how far from stability the initial state of the disk was chosen. The start from different initial disk thicknesses led to models without any systematic differences between them. Figure 9 illustrates this result, which is similar to that obtained in their numerical simulations by Sellwood and Merritt (1994). Thus, our three-dimensional calculations are in good agreement with the conclusion following from our analysis of the dispersion relation for a thin disk that a bulge is an effective stabilizing factor during the growth of bending instability. Moreover, since the initial bend is formed in the most unstable central part of the galaxy (Sotnikova and Rodionov 2003), it is the central regions that must be stabilized. This does not require a massive dark halo; the presence of a compact spherical component like a bulge will suffice. This suggests that the final thickness of the model galaxy depends not only on the total 14 mass of the spherical component, but also on the mass distribution in it. Barred galaxies. The bending instability of bars is an effective vertical disk heat- ing mechanism for galaxies unstable against the formation of a bar (Raha et al. 1991; Sotnikova and Rodionov 2003). In contrast to the warp in the entire disk, the warp in the bar is formed not in the central regions, but in the entire bar simultaneously (this is seen particularly clearly in our color two-dimensional histograms of the warp ac- cessible at http://www.astro.spbu.ru/staff/seger/articles/warps 2002/fig6 web.html and http://www.astro.spbu.ru/staff/seger/articles/warps 2002/fig7 web.html). Therefore, the conclusion that a compact bulge during the growth of bending instability in a bar will have the same effective stabilizing effect as that for hot stellar disks is not obvious in advance. In our simulations with QT = 1.5 and µ <∼1.0, the warp in the bar was formed early, at t ≈ 800. The presence of a compact bulge eventually led to the formation of thinner disks, although the effect itself was fairly complex. Within R < 1.5h, the most prominent features of the bars at late evolutionary stages were X-shaped structures. If the disk is viewed edge-on, then they manifest themselves as a bulge with an appreciable extent in the z direction with boxy isophotes. In the region R < 1.5h where the bar dominated, we failed to reveal any distinct patterns in the model disk thickness variations with increasing contribution of the bulge to the total mass of the spheroidal component. In general, the presence of a bulge "pushed forward" the bar formation time and caused the saturation level for the bending instability of a bar to lower. However, a further analysis is required to completely understand the processes during the interaction between a compact bulge and a bar. As regards the peripheral regions of the disks (R > 1.5h), they were always apprecia- bly thinner in the models with a compact bulge at late evolutionary stages. Figure 10 demonstrates this effect for two groups of models with different bulge contributions: mod- els 52 and 51 with µ = 0.5 and models 43 and 42 with µ = 0.875. The differences in thickness show up most clearly in the models with a small relative mass of the spheroidal component. When µ increases, the disk becomes very thin, as might be expected, and its thickness ceases to depend on how the mass is distributed between the halo and the bulge. 6 CONCLUSIONS A comparison of the conclusions that follow from a linear analysis with the results of numerical simulations for three-dimensional disks shows that, in contrast to homogeneous models, global bending modes with a wavelength longer than the disk scale length can arise in inhomogeneous disks. If the amplitude of the waves during the growth of instability increases significantly, then they heat it up significantly in the vertical direction as they pass through the entire disk. Hot disks are most unstable against the growth of bending perturbations. An additional spheroidal component, for example, a dark halo is a factor that stabilizes the bending perturbations. Our additional qualitative analysis of the dispersion relation for inhomogeneous models led us to new conclusions regarding the stabilization conditions for the bending modes in stellar disks. These conclusions were confirmed in our numerical simulations. 15 (1) Since the central regions of the disk (particularly if the disk is hot) are most unsta- ble, the conditions under which the growth of perturbations is suppressed are determined not only by the mass of the spherical component, but also by the density distribution in it. The suppressing effect is enhanced with increasing concentration toward the center. (2) The presence of a compact and moderately massive bulge in a galaxy effectively prevents the growth of bending perturbations. (3) It follows from an analysis of the entire set of our results that a more accurate approach to estimating the dark halo mass from the observed relative thickness of the stellar disk z0/h in spiral galaxies is required. ACKNOWLEDGMENTS This work was supported by the Russian Foundation for Basic Research (project no. 03-02-1752), the Federal "Astronomy" Program (project no. 40.022.1.1.1101), and a grant from the President of Russia for support of leading scientific schools (NSh-1088.2003.2). REFERENCES 1. S. Araki, Ph. D. Thesis, Massachus. Inst. Tech. (1985). 2. J. Barnes and P. Hut, Nature 324, 446 (1986). 3. J. Binney and S. Tremaine, Galactic Dynamics (Princeton Univ. Press, Princeton, 1987). 4. E. Griv, M. Gedalin, and Chi Yuan, Astrophys. J. 580, L27 (2002). 5. E. Griv and Tzihong Chiueh, Astrophys. J. 503, 186 (1998). 6. L. Hernquist, Astrophys. J., Suppl. Ser. 86, 389 (1993). 7. P. C. van der Kruit and L. Searle, Astron. Astrophys. 95, 105 (1981). 8. R. M. Kulsrud, J.W.-K. Mark, and A. Caruso, Astrophys. Space Sci. 14, 52 (1971). 9. D. Merritt, Astron. J. 111, 2462 (1996). 10. D. Merritt and J. A. Sellwood, Astrophys. J. 425, 551 (1994). 11. E. A. Mikhailova, A. V. Khoperskov, and S. S. Sharpak, Stellar Dynamics -- from Clasic to Modern Ed. by L. P. Ossipkov and I. I. Nikiforov (St. Petersburg State Univ. Press, St. Petersburg, 2001), p. 147. 12. P. A. Patsis, E. Athanassoula, P. Grosbol, and Ch. Skokos Ch., Mon. Not. R. Astron.Soc. 335, 1049 (2002). 13. V. L. Polyachenko and I. Sh. Shukhman, Pis'ma Astron. Zh. 3, 254 (1977) [Sov. Astron. Lett. 3, 134 (1977)]. 14. N. Raha, J. A. Sellwood, R. A. James, and F.D. Kahn, Nature 352, 411 (1991). 15. J. A. Sellwood, Astrophys. J. 473, 733 (1996). 16. J. A. Sellwood and R. G. Carberg, Astrophys. J. 282, 61 (1984). 17. J. A. Sellwood and D. Merritt, Astrophys. J. 425, 530 (1994). 18. N. Ya. Sotnikova and S. A. Rodionov, Pis'ma Astron. Zh. 29, 367 (2003) [Astron. Lett. 29, 321 (2003)]. 19. L. Spitzer, Astrophys. J. 95, 325 (1942). 20. L. Spitzer and M. Schwarzshild, Astrophys. J. 114, 385 (1951). 21. L. Spitzer and M. Schwarzshild, Astrophys. J. 118, 106 (1953). 22. P. J. Teuben, ASP Conf. Ser. 77, 398 (1995). 23. A. Toomre, Astrophys. J. 139, 1217 (1964). 16 24. A. Toomre, Geophys. Fluid Dyn., N66-46, 111 (1966). 25. Yao-Huan Tseng, Chinese Journ. Phys. 38, 111 (2000). 26. H. Velasquez and S. D. M. White, Mon. Not. R. Astron. Soc. 304, 254 (1999). 27. I. W. Walker, J. Ch. Mihos, and L. Hernquist, Astrophys. J. 460, 121 (1999). 28. A. V. Zasov, D. V. Bizyaev, D. I. Makarov, and N. V. Tyurina, Pis'ma Astron. Zh. 28, 599 (2002) [Astron. Lett. 28, 527 (2002)]. 29. A. V. Zasov, D. I. Makarov, and E. A. Mikhailova, Pis'ma Astron. Zh. 17, 884 (1991) [Sov. Astron. Lett. 17, 374 (1991)]. Translated by V. Astakhov 17 Fig. 1: (a) Magnitude of the vertical gravitational force (−Fz) versus z at various distances R from the disk center; (b) the square of the vertical oscillation frequency ν2 d = −∂Fz/∂z versus z for various R. We took G = 1 and the disk parameters Md = 1 (the total disk mass), h = 3.5, and z0 = 0.1. 18 Fig. 2: Ratio σz/σR versus R for the time t = 3000 (model 12 with a massive halo: µ = 3.0, µb = 0). 19 0.4 0.35 0.3 0.25 c V 0.2 0.15 0.1 0.05 0 0 Model 50 Model 76 Model 75 Model 49 Model 53 2 4 6 8 R 10 12 14 Fig. 3: Initial rotation curves for models 50, 76, 75, 49, and 53. The ratio of the total mass of the spherical component to the mass of the disk within a radius of 4h is the same for all models, µ = 0.5; µb = 0 for model 50, µb = 0.0625 for model 76, µb = 0.125 for model 75, µb = 0.25 for model 49, and µb = 0.5 for model 53. The unit of velocity is 978 km s−1. 20 Fig. 4: Evolution of the velocity dispersions σR and σz at R = 2h for the same models as those in Fig. 3. 21 Fig. 5: Edge-on view of the galaxy at the time t = 3000 for models 50, 76, 75, 49, and 53. The strenght of the image blackening corresponds to the logarithm of the particle number per pixel. The horizontal and vertical scales are 60 and 10, respectively. The ratio of the total mass of the spherical component to the mass of the disk within a radius of 4h is the same for all models, µ = 0.5. 22 Fig. 6: Radial thickness profiles for the galaxy at the time t = 3000 for model 53. The thickness was determined by two methods: as the rms value of the z coordinates of the disk particles, zrms, and as twice the median of z, 2z1/2. 23 Fig. 7: Radial thickness profiles for the galaxy (2z1/2) at the time t = 3000 for models 50, 76, 75, 49, and 53. 24 Fig. 8: Evolution of the disk thickness (2z1/2) at R = 2h for models 50, 76, 75, 49, and 53 (see the caption to Fig. 3). 25 Fig. 9: Radial thickness profiles for the galaxy (2z1/2) at the time t = 3000 for models that differ only by the initial thickness. For all of the models, µ = 0.6 and µb = 0. Models 11 1, 11 2, and 11 3 are different random realizations of a stellar system with z0 = 0.1; models 25 and 25 1 are different random realizations of a system with z0 = 0.2; model 26 1 is for z0 = 0.3. 26 Fig. 10: Radial thickness profile for the galaxy (2z1/2) at the time t = 3000 for models unstable against the growth of a bar mode (QT = 1.5): (a) models with µ = 0.5 (model 52 with µb = 0, model 51 with µb = 0.25), (b) models with µ = 0.875 (model 43 with µb = 0.25, model 42 with µb = 0.25); the initial half-thickness for all of the models is z0 = 0.1. 27
0811.4173
1
0811
2008-11-25T21:00:00
Parameter degeneracies and (un)predictability of gravitational microlensing events
[ "astro-ph" ]
(abridged) Some difficulties in determining the physical properties that lead to observed anomalies in microlensing light curves, such as the mass and separation of extra-solar planets orbiting the lens star, or the relative source-lens parallax, are already anchored in factors that limit the amount of information available from ordinary events and in the adopted parametrization. Moreover, a real-time detection of deviations from an ordinary light curve while these are still in progress can only be done against a known model of the latter, and such is also required for properly prioritizing ongoing events for monitoring in order to maximize scientific returns. Despite the fact that ordinary microlensing light curves are described by an analytic function that only involves a handful of parameters, modelling these is far less trivial than one might be tempted to think. A well-known degeneracy for small impacts, and another one for the initial rise of an event, makes an interprediction of different phases impossible, while determining a complete set of model parameters requires the assessment of the fundamental characteristics of all these phases. While the wing of the light curve provides valuable information about the time-scale that absorbs the physical properties, the peak flux of the event can be meaningfully predicted only after about a third of the total magnification has been reached. Parametrizations based on observable features not only ease modelling by bringing the covariance matrix close to diagonal form, but also allow good predictions of the measured flux without the need to determine all parameters accurately. Campaigns intending to infer planet populations from observed microlensing events need to invest some time into acquiring data that allows to properly determine the magnification function.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000–000 (0000) Printed 29 May 2018 (MN LATEX style file v2.2) Parameter degeneracies and (un)predictability of gravitational microlensing events M. Dominik,1 ⋆† 1SUPA, University of St Andrews, School of Physics & Astronomy, North Haugh, St Andrews, KY16 9SS, United Kingdom 29 May 2018 ABSTRACT Some of the difficulties in determining the underlying physical properties that are relevant for observed anomalies in microlensing light curves, such as the mass and separation of extra- solar planets orbiting the lens star, or the relative source-lens parallax, are already anchored in factors that limit the amount of information available from ordinary microlensing events and in the way these are being parametrized. Moreover, a real-time detection of deviations from an ordinary light curve while these are still in progress can only be done against a known model of the latter, and such is also required for properly prioritizing ongoing events for monitor- ing in order to maximize scientific returns. Despite the fact that ordinary microlensing light curves are described by an analytic function that only involves a handful of parameters, mod- elling these is far less trivial than one might be tempted to think. A well-known degeneracy for small impacts, and another one for the initial rise of an event, makes an interprediction of different phases impossible, while in order to determine a complete set of model parameters, the fundamental characteristics of all these phases need to be properly assessed. While it is found that the wing of the light curve provides valuable information about the time-scale that absorbs the physical properties, the peak flux of the event can be meaningfully predicted only after about a third of the total magnification has been reached. Parametrizations based on ob- servable features not only ease modelling by bringing the covariance matrix close to diagonal form, but also allow good predictions of the measured flux without the need to determine all parameters accurately. Campaigns intending to infer planet populations from observed mi- crolensing events need to invest some fraction of the available time into acquiring data that allows to properly determine the magnification function. Key words: gravitational lensing – planetary systems. 8 0 0 2 v o N 5 2 ] h p - o r t s a [ 1 v 3 7 1 4 . 1 1 8 0 : v i X r a 1 INTRODUCTION An efficient detection of planets by means of gravitational mi- crolensing requires sufficiently accurate predictions of the under- lying ordinary light curve against which the planetary deviations need to be identified (Dominik et al. 2007, 2008). An optimal mon- itoring strategy in order to maximize the scientific return moreover profits strongly from the proper determination of a full set of model parameters as early as possible, and the latter becomes a require- ment for finally assessing the planet detection efficiency and draw- ing conclusions about the planet population. It is known that early-stage event prediction suffers from degeneracies (Albrow 2004), in particular the peak flux is hard to assess, while the event time-scale tE, required to relate the observations to the underlying physical properties, can remain strongly uncertain even after the event has been observed over its full course, but not covered well enough (Wo´zniak & Paczy´nski ⋆ Royal Society University Research Fellow † E-mail: [email protected] 1997). The lack of determinacy of tE is particularly apparent for strongly-blended events, which in fact comprise the full sample for observations towards neighbouring galaxies such as M31 (e.g. Baillon et al. 1993). Understanding of ordinary light curves (comprising single point-like source and lens stars), their optimal parametrization, and the apparent degeneracies and ambiguities, is also crucial and use- ful for modelling events that involve anomalies, and drawing con- clusions about e.g. stellar binaries, stellar masses derived from par- allax measurements, and planets. Sparse event coverage in criti- cal regions is even prone to lead to ordinary events allowing for multiple minima of the χ2 (least-squares) hypersurface, this fact being further complicated by the presence of potential outliers in the data and the application of robust-estimation techniques to deal with these (e.g. Dominik et al. 2007). In this paper, different phases of ordinary microlensing events along with their characteristics are identified, and it is shown how feature-oriented parametrizations can be used to quantify the be- haviour while avoiding parameter correlations. Moreover, some fundamental requirements for a monitoring strategy that allows 2 M. Dominik to meet the goal of determining either the observed flux or the corresponding magnification are discussed. The basis for this is formed by approximate relations between the magnification and the angular separation between lens and source that make the light curve independent of the impact parameter. Together with some arising degeneracies, these have already been identified by Wo´zniak & Paczy´nski (1997), but here some new light is shed on the implications and the focus is on different aspects that are emer- gent right now, which leads to arriving at some new and different conclusions. While Sect. 2 reviews the general properties of ordinary mi- crolensing events and the canonical parametrization, and Sect. 3 presents the advantages of an alternative parametrization that is oriented towards the observable characteristic features, Sect. 4 dis- cusses the predictability of ordinary microlensing events or the lack of it. A summary and final conclusions are provided in Sect. 5. 2 THE CANONICAL TREATMENT OF ORDINARY EVENTS A microlensing event results if two stars happen to be closely aligned as seen from Earth, where this alignment is quantified by the unique characteristic scale of gravitational microlensing, namely the angular Einstein radius (Einstein 1936) event time-scale tE. It is in fact its relation to the physical event properties that makes the parameter tE a preferred choice amongst possible time-scales. As illustrated in Fig. 1, ordinary light curves: (1) are symmetric with respect to a peak at epoch t0, (2) reach a peak flux there, (3) approach a baseline flux for times far away from the peak, and (4) show characteristic inflection points. How- ever, the parameter t0 is the only one that directly relates to the characteristic features of the light curve, while all the others are not a proper reflection. This is not at all favourable for modelling, and instead being able to essentially read off the model parameters from the collected data would ease life a lot. The phenomenon of gravitational microlensing leads to a char- acteristic magnification A(t; p) as a function of time t and the model parameters p that describe the lens-observer-source geome- try, the lens properties, and the source brightness profile. The ob- served flux F (k)(t; p) for a given site and passband - denoted by the multi-index k -, however, furthermore is a linear function of the intrinsic flux of the observed source star F (k) S , which is magni- fied, and a background flux F (k) B , where (e.g. Albrow et al. 2000) F (k)(t; p) = F (k) S A(t; p) + F (k) B . (5) This allows us to isolate these two parameters from the remaining parameter space, so that θE =r 4GM c2 πLS 1 AU , (1) p = (F (k) S , F (k) B , p) , and for every value of p, minimizing (6) (7) where M denotes the mass of the foreground lens star, G is the universal gravitational constant, c is the vacuum speed of light, and πLS = 1 AU (cid:0)D−1 L − D−1 S (cid:1) stands for the relative source-lens parallax, with DL and DS be- ing the distance from Earth to the foreground (lens) star and the observed background source star, respectively. With u θE denoting the angular separation between lens and source star, gravitational bending of light by the lens star yields an observable magnification of the source star as an analytic function of u, which reads (2) A(u) = u2 + 2 u √u2 + 4 . (3) Stellar kinematics implies a non-vanishing relative proper mo- tion µ between lens and source star, so that u is a function of time and the magnification A(u) describes a characteristic light curve. While in the earlier history of gravitational microlensing several different event time-scales have been used (e.g. Griest 1991; de R´ujula et al. 1991), tE ≡ θE/µ has emerged as a popular con- venient choice. For uniform proper motion, i.e. constant µ, the dimensionless separation u can be expressed by means of three quantities that form the parameter vector p = (u0, t0, tE), so that , (4) u(t; u0, t0, tE) =ru2 tE (cid:17)2 0 +(cid:16) t − t0 where u0 θE is the smallest angular separation, encountered at epoch t0, while the source moves by an angular Einstein radius relative to the lens within the time-scale tE (Paczy´nski 1986). Whereas u0 and t0 only define the position of the trajectory of the source relative to the lens, and therefore do not carry any informa- tion about the relevant physical properties that determine the mag- nification, which are the relative parallax πLS, the relative proper motion µ, and the lens mass M, all these are convolved into the χ2(t(k) i ; p) = s Xk=1 i nk Xi=1 F (t(k) ; p) − F (k) σ(k) Fi i !2 , thereby obtaining a maximum-likelihood estimate of the parameter vector p, means that the best-fitting source and background fluxes can be expressed in closed analytical form (Rattenbury 2003) σ2 σ2 σ2 σ2 FS = P A(ti)Fi Fi P 1 P [A(ti)]2 Fi P 1 FB = P [A(ti)]2 Fi P Fi P [A(ti)]2 Fi P 1 σ2 σ2 σ2 , σ2 σ2 σ2 σ2 Fi σ2 Fi Fi −P A(ti) Fi P Fi Fi (cid:19)2 Fi −(cid:18)P A(ti) Fi P A(ti)Fi Fi −P A(ti) Fi (cid:19)2 Fi −(cid:18)P A(ti) Xi=1 A(t(k) nk σ2 σ2 s i Xk=1 while the non-linear minimization process can be restricted to , (8) χ2(t(k) i , F (k) i , σ(k) Fi ; p) = ; p) − A(k) σ(k) Ai i !2 , (9) expressed by means of the magnification rather than the flux, where A(k) i = and σ(k) Ai = i − F (k) F (k) F (k) B S . σ(k) Fi F (k) S (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) (10) (11) (Un)predictability of microlensing events 3 Figure 1. Light curves for a selection of impact parameters u0, where the event magnification A = (F − FB)/FS is plotted as a function of (t − t0)/tE, where the closest angular approach u0 is realized at t0, tE = θE/µ, with θE denoting the angular Einstein radius, defined in Eq. (1), and µ being the relative proper motion between lens and source star. Figure 2. Light curves for the same impact parameters as adopted for Fig. 1, but now the relative flux difference (F − Fbase)/∆F with respect to its peak is plotted as a function of (t − t0)/t1/2, where the half-maximum time-scale t1/2 is defined by Eq. (15). All light curves range between the asymptotics for u ≪ 1 and u ≫ 1 (shown in black), as given by Eq. (19). 3 PARAMETERS THAT MATCH OBSERVATIONAL FEATURES It is well known that avoiding parameter degeneracies and strong correlations eases the modelling process. Moreover, a careful study of these provides valuable insight into the structure of parame- ter space, which can be used as a guide for developing observ- ing strategies and explicitly shows their limitations. For ordinary microlensing light curves, it turns out that the observed flux can be rewritten by means of a different parametrization that better matches the observational features, which are key to diagonalizing the covariance matrix. With the baseline flux F (k) base and the peak flux F (k) 0 given by (12) (13) (14) base and ∆F (k), the ob- base = F (k) F (k) F (k) = F (k) S + F (k) B , S A[u(t0; p)] + F (k) B , 0 the maximal flux difference ∆F (k) reads ∆F (k) = F (k) base = F (k) 0 − F (k) S [A(u0) − 1] . B , but also in F (k) S and F (k) Not only in F (k) served flux F (k)(t; p) is a linear function, namely F (k)(t; p) = ∆F (k) A[u(t; p)] − 1 A(u0) − 1 + F (k) base , 1 2 base = ∆F (k) . so that these parameters can be separated by a linear fit as well. One can then define a half-maximum time t1/2, so that at time t0± t1/2, half of the flux offset is encountered, i.e. F (k)(t0 ± t1/2; u0, t0, tE) − F (k) This means that the epochs t0 ± t1/2 correspond to a magnification (16) A1/2(u0) ≡ A[u(t0 ± t1/2; u0, t0, tE)] = With u1/2 ≡ u(t0 ± t1/2; u0, t0, tE) and by means of Eq. (4), one finds t1/2 = tEqu2 1/2 − u2 0 , A(u0) + 1 (15) (17) 2 . which leads to . (18) 1/2 u2 u(t; u0, t0, t1/2) = u0s1 +(cid:18) u2 t1/2 (cid:19)2 0 − 1(cid:19) (cid:18) t − t0 If one now plots the offset flux F (k)(t; p) − F (k) base in units of the difference ∆F (k) between peak and baseline, and scales t − t0 with t1/2 rather than tE, as illustrated in Fig. 2, one finds that light curves with different u0 nearly coincide around their peaks. In con- trast, it is the wing region of the event 1.5 t1/2 . t − t0 . 3 t1/2 that is best-suited to provide information about the impact param- eter u0, and thereby of the event time-scale tE, related to the un- derlying physical properties.1 In principle, u0 is determined by the slope of the light curve at t = t0 ± t1/2, but the narrow range2 strongly limits the feasibility of such an approach in practice. It is also instructive to explore the limits of small and large separations more closely. Both for u ≪ 1 and u ≫ 1, the observed relative flux offset (F (k)(t)−F (k) base)/∆F (k) becomes independent of u0, while t1/2 becomes proportional to u0 tE, however with dif- ferent proportionality factors for the two extreme cases. As summa- rized in Table 1, this is due to the fact that the magnification can rea- sonably well be approximated by a more simple expression, namely [A(u)−1]/[A(u0)−1] approaching u0/u for u ≪ 1, and (u0/u)4 for u ≫ 1. For u ≪ 1, one retrieves the known results that apply to microlensing of unresolved sources where F (k) S , being an inevitability for observations towards M31 or other nearby other galaxies (Baillon et al. 1993; Wo´zniak & Paczy´nski 1997). Look- ing at Fig. 2 again, one sees that all light curves range between the two extreme cases B ≫ F (k) F (k)(t) − F (k) base ∆F (k) ≃ t1/2(cid:17)2(cid:21)−2 (cid:20)1 + (√2 − 1) (cid:16) t−t0 t1/2(cid:17)2(cid:21)−1/2 (cid:20)1 + 3 (cid:16) t−t0   (u ≫ 1) , (19) (u ≪ 1) . The wings of the light curves converge towards the expression for 1 The subsequent section shows that the asymptotic degeneracy for large u does not extend into this region. 2 The absolute value of the slope ranges between 0.375 (for u0 → 0) and 2 − √2 ≈ 0.586 (for u0 → ∞). 4 M. Dominik Table 1. Asymptotic behaviour of the observed flux F (k), as well as of other relevant quantities, in the limits u ≪ 1 or u ≫ 1, and its matching parametriza- tion. A(u) A[u(t; p)] − 1 A(u0) − 1 A1/2 u1/2 t1/2 u(t) F (k)(t; ∆F (k), F (k) base, t0, t1/2) − F (k) base ≃ u ≪ 1 u−1 u0 u(t) 1 2 u0 2 u0 √3 u0 tE ≃ ≃ ≃ ≃ ≃ ≃ u0s1 + 3(cid:18) t − t0 t1/2 (cid:19)2 s1 + 3(cid:18) t − t0 t1/2 (cid:19)2 ∆F (k) u ≫ 1 1 + 2 u−4 u(t)(cid:17)4 (cid:16) u0 1 + 1 u4 0 4√2 u0 (√2 − 1) u0 tE u0s1 + (√2 − 1)(cid:18) t − t0 t1/2 (cid:19)2 t1/2 (cid:19)2#2 "1 + (√2 − 1)(cid:18) t − t0 ∆F (k) u ≫ 1, where for smaller u0, the transition from the u ≪ 1 to the u ≫ 1 asymptotic occurs at larger t − t0/t1/2. A(u) ≃ 1 + 2 (cid:18) tE t − t0(cid:19)4 (u ≫ u0, u ≫ 1) , (20)   tE t − t0 (u0 ≪ u ≪ 1) , where the second case is realized for a substantial time interval if u0 is sufficiently small. With the blend ratio g(k) = F (k) B /F (k) S , the observed flux F (k)(t) can be written as 4 DIFFERENT EVENT PHASES AND THE LACK OF PREDICTABILITY One might think that 3 parameters (like u0, t0, and tE) can be ob- tained straightforwardly by means of regression from as few as 4 data points, but such an attempt fails if the observable does not sig- nificantly change with a variation of the considered parameter. So far, it has been assumed that the light curve has been sampled over its full course, so that a full set of characteristics can be determined from which model parameters can be derived. However, the light curve develops in time, so that some characteristics are not accessi- ble at early stages. While the baseline flux F (k) base is being observed much before any rise in brightness occurs, the peak flux F (k) or the flux shift ∆F (k) = F (k) base, respectively, remain unknown, as we shall see more explicitly in the following. 0 − F (k) 0 Reviewing the asymptotics for u ≪ 1 and u ≫ 1 from a different perspective already shows us that there is no interpre- dictability between the peak and wing regions of the observed light curve, and in particular, early observations give poor estimates of the peak magnification, as well as on the time-scale tE. In partic- ular, maximum-likehihood estimates corresponding to values that minimize χ2, Eq. (9), frequently yield very small u0, far away from expectations. Therefore, Albrow (2004) has suggested to use a maximum-a-posteriori estimate instead, incorporating the actual distribution of the parameters of the observed events of the mi- crolensing surveys as prior. Let us look into this with a view on the parameter degen- eracies. The earliest stages of an event are characterized by t − t0/tE ≫ u0, so that u ≃ t − t0/tE. For the two extreme cases of u being far from unity, one finds 1 + g(k) 1 + g(k) base (cid:20) A[u(t; p)] + g(k) (cid:21) + 1(cid:21) , base (cid:20) A[u(t; p)] − 1 t − t0(cid:19)4 1 + g(k) (cid:18) tE 2 (u ≫ u0, u ≫ 1) , 1 tE 1 + g(k) t − t0 = t(k) s t − t0 (u0 ≪ u ≪ 1) , , t(k) s ≡ tE 1 + g(k) , F (k)(t) = F (k) = F (k) so that F (k)(t) F (k) base where t(k) r ≡ −1 ≃   4p1 + g(k) tE = 2 (cid:18) t(k) t − t0(cid:19)4 r are characteristic rise times, absorbing tE and g(k), with tr = ts = tE for g = 0. Figure 3 shows F (k)(t)/F (k) base − 1 as a function of (t− t0)/tr in a double-logarithmic plot, so that the approximate relations of Eq. (22) correspond to straight lines. It illustrates that different be- haviour allows to distinguish three phases of a microlensing event, corresponding to the initial rise, a mid-phase, and the peak ap- (21) (22) (23) (Un)predictability of microlensing events 5 0 /F (k) Figure 4. Development of events from the peak with a given relative bright- ening F (k) base, where light curves corresponding to three selected val- ues are shown, towards the baseline flux F (k) base for two different blend ratios g ≡ F (k) B . Given that the peak region is characterized by its width, F (k) base − 1 is shown as function of (t − t0)/t1/2 (in a double-logarithmic plot). As for Figure 3, the asymptotic behaviour given by Eq. (22) is shown by means of black lines. S /F (k) 0 /F (k) can be made to match well near the peak, whereas observations in the wing of the light curve on the departure from the asymptotic mid-phase behaviour (u0 ≪ u ≪ 1) or on the transition into the baseline-approach phase (u ≫ 1) are useful to determine tE. The rather long duration of these phases favours such an attempts by not requiring a high sampling rate for obtaining a larger number of measurements. 5 SUMMARY AND FINAL CONCLUSIONS Despite the fact that the timescale tE ≡ θE/µ, where θE de- notes the angular Einstein radius, and µ the relative proper mo- tion between lens and source star, is the crucial one for drawing conclusions about the underlying physical properties that led to a microlensing event, it is not a good choice for describing the ob- servable characteristic features, and thereby not a useful means of describing or predicting events in progress. Contrary to common belief, 3 model parameters cannot al- ways properly be extracted from a least-squares fit involving at least 4 data points. In fact, such attempts fail if the respective function to be matched to the observed data does not significantly depend on each of the parameters over the region where data have been acquired. Microlensing light curves usually go through 3 phases from baseline to peak as well as from peak back to baseline: Two rise phases characterized by different rise times t(k) , corre- sponding to different power laws of the magnification with the an- gular separation between lens and source, as well as a peak region, characterized by a half-width t1/2. None of the individual phases contains characteristic information about the complete set of model parameters, and in order to reveal accurate estimates for all of them, each of them requires appropriate coverage. While the mid-phase and t(k) s r S B /F (k) Figure 3. Development of a microlensing event from a common initial rise to different peak fluxes for three selected impact parameters u0 (colour- coded) and two blend ratios g(k) ≡ F (k) (solid or dashed line). Double-logarithmic plot showing the relative brightening F (k)(t)/F (k) base− 1 above baseline as a function of (t − t0)/t(k) is the rise time defined by Eq. (23), so that for the two selected blend ratios tE = t(k) (g(k) = 0) or tE = 2 t(k) (g(k) = 15), respectively. The black lines correspond to the asymptotic behaviour given by Eq. (22). Be aware of the fact that the same scale corresponds to tiny changes in the observed flux near the bottom of the plot, but huge ones near the top. Similarly, the left parts span large time intervals, while the right parts span small ones. , where t(k) r r r r proach, where a determination of the full model parameter set re- quires an assessment of the fundamental characteristics of all these phases. With the approximate F (k)(t) diverging as t → t0, an esti- mate for t0 based on such an approximation corresponds to u0 = 0. Only a departure from the approximation in the form of an evident turn-off gives evidence for non-vanishing u0. However, as Fig. 3 illustrates, this happens at rather late stage, and roughly when a third of the peak magnification has been reached. Given that for u . 1, A(u) differs substantially from the asymptotic behaviour for u ≫ 1, the blend ratio g(k) can in principle be determined rather early, which then provides the time-scale tE that is related to the underlying physical properties of the event. In practice this re- quires sufficiently dense and precise measurements, which are fre- quently not available for fluxes close to Fbase, but the long duration of the rise phase in principle allows for lots of data to be collected. Nevertheless, as soon as both t(k) can be determined from the acquired data, the blend ratio g(k) and tE are known. Again, one sees the power of observations covering the wing of the light curve. and t(k) s r Going the other way round and coming from the peak, the microlensing light curve first follows the decay with t1/2, then en- ters the mid-phase characterized by t(k) , and finally follows a de- crease described by t(k) . Double-logarithmic plots of the relative r offset brightening F (k) base − 1 as a function of (t − t0)/t1/2 are shown in Fig. 4, while the relevant magnifications, impact pa- rameters, and time-scales for the selected cases are listed in Table 2. Again, one sees that the light curves for different blend ratio and tE 0 /F (k) s 6 M. Dominik Table 2. Magnifications, impact parameters and time-scales for the relative peak fluxes adopted in Fig. 4. 0 /F (k) F (k) base A0 A1/2 u0 u1/2 tE/t1/2 t(k) r /t1/2 t(k) s /t1/2 50 20 5 785 305 65 25.5 10.5 3 393 153 33 0.0200 0.0500 0.203 g = 0 0.0392 0.0956 0.348 g = 15 0.00127 0.00328 0.0154 0.00255 0.00654 0.0303 29.6 12.3 3.53 453 177 38.3 50 20 5 50 20 5 29.6 12.3 3.53 227 88.4 19.1 29.6 12.3 3.53 28.4 11.1 2.39 tE = 2 t(k) r = 16 t(k) s for g = 15, while tE = t(k) r = t(k) s for g = 0. REFERENCES Albrow M. D., et al., 2000, ApJ, 535, 176 Albrow M. D., 2004, ApJ, 607, 821 Baillon P., Bouquet A., Giraud-Heraud Y., Kaplan J., 1993, A&A, 277, 1 de R´ujula A., Jetzer P., Mass´o E., 1991, MNRAS, 250, 348 Dominik M., et al., 2007, MNRAS, 380, 792 Dominik M., et al., 2008, AN, 329, 248 Einstein A., 1936, Science, 84, 506 Gaudi B. S., et al., 2002, ApJ, 566, 463 Griest K., 1991, ApJ, 366, 412 Han C., 2007, ApJ, 661, 1202 Paczy´nski B., 1986, ApJ, 304, 1 Rattenbury N. J., 2003, PhD thesis, University of Auckland Snodgrass C., Tsapras Y., Street R., Bramich D., Horne K., Do- minik M., Allan A., 2008, PoS(GMC8)056 Wo´zniak P., Paczy´nski B., 1997, ApJ, 487, 55 gets squeezed for impact angles u θE of the order of the angular Einstein radius θE or larger, the part of the light curve with u ≫ 1 becomes indistinguishable from the baseline for strongly-blended events, so that on the approach to baseline, the offset magnification is proportional to u−1 rather than u−4, given that still u ≪ 1 for t − t0 ≫ 1. It is well-known that an appropriate estimate of event pa- rameters at early event stages is not feasible, and in particular the peak magnification is regularly overpredicted by a maximum- likelihood estimate corresponding to minimizing the sum of nor- malized squared deviations χ2. Just for this reason, Albrow (2004) had suggested to use a maximum-a-posteriori estimate instead, with a suitable prior. While this brings the estimate closer to its expecta- tion value, it does not get around the uncertainty. A closer exima- tion shows that the light curve is compatible with an infinite peak flux until roughly a third of the true offset magnification is reached. Moreover, it is a wing region 1.5 t1/2 . t − t0 . 3 t1/2 that is best suited to determine the blend ratio g(k) = F (k) (and with it the time-scale tE, rather than the immediate vicinity of the peak. The rather long duration of this phase allows to obtain a suit- able measurement without the need for very dense sampling. B /F (k) S For an accurate prediction of the observed flux, a proper deter- mination of the full set of model parameters is not required, so that local approximations can provide a reasonable substitute. In sharp contrast, proper knowledge of tE, which implies knowledge of the blend ratio g(k) and the magnification A(t), is a requirement for determining the event detection efficiency to planets (Gaudi et al. 2002) as well as for prioritising ongoing events in order to maxi- mize it (Han 2007; Snodgrass et al. 2008). Without such informa- tion, one neither knows the amplitude, nor the duration, nor the location of potentially arising planetary signals. Therefore, an ef- ficient campaign for inferring the planet population from observed microlensing events needs to invest time into observations that al- low to properly determine the event parameters, rather than just try- ing to detect planets in poorly determined events, where unsuitable assumptions about model parameters may yield to bad choices, or efforts could even turn out to be wasted if the planet detection ef- ficiency cannot be assessed. Building upon the findings presented in this paper, a more detailed study of event (un)predictability tak- ing into account the specific capabilities of observing campaigns could hence provide important clues towards optimizing strategies for detecting planets and determining their population statistics.
astro-ph/0403227
1
0403
2004-03-09T23:12:27
The Power of Exploratory Chandra Observations
[ "astro-ph" ]
With its excellent spatial resolution, low background, and hard-band response, the Chandra X-ray Observatory is ideal for performing exploratory surveys. These efficient, sensitive observations can place constraints on fundamental properties of a quasar continuum including the X-ray luminosity, the ratio of X-ray to UV power, and the X-ray spectral shape. To demonstrate the power of such surveys to provide significant insight, we consider two examples, a Large Bright Quasar Survey sample of broad absorption line quasars and a sample of Sloan Digital Sky Survey (SDSS) quasars with extreme CIV blueshifts. In both cases, exploratory Chandra observations provide important information for a physical understanding of UV spectroscopic differences in quasars.
astro-ph
astro-ph
AGN Physics with the Sloan Digital Sky Survey ASP Conference Series, Vol. 311, 2004 G.T. Richards and P.B. Hall, eds. The Power of Exploratory Chandra Observations Sarah C. Gallagher University of California, Los Angeles, Division of Astronomy & Astrophysics, 405 Hilgard Avenue, Los Angeles, CA 90095 Gordon T. Richards Princeton University Observatory, Peyton Hall, Princeton, NJ 08544 W. Nielsen Brandt & George Chartas The Pennsylvania State University, Department of Astronomy & Astrophysics, 525 Davey Laboratory, University Park, PA 16802 Abstract. With its excellent spatial resolution, low background, and hard-band response, the Chandra X-ray Observatory is ideal for perform- ing exploratory surveys. These efficient, sensitive observations can place constraints on fundamental properties of a quasar continuum including the X-ray luminosity, the ratio of X-ray to UV power, and the X-ray spectral shape. To demonstrate the power of such surveys to provide sig- nificant insight, we consider two examples, a Large Bright Quasar Survey sample of broad absorption line quasars and a sample of Sloan Digital Sky Survey (SDSS) quasars with extreme C iv blueshifts. In both cases, exploratory Chandra observations provide important information for a physical understanding of UV spectroscopic differences in quasars. 1. Introduction X-ray emission appears to be a universal signature of quasar spectral energy distributions, confirming expectations from accretion physics. Based on rapid variability of soft X-rays in conjunction with the standard black-hole paradigm, these photons are believed to be emitted from the region immediately surround- ing the black hole. Energetically, X-rays are significant, contributing 2 -- 20% of the bolometric luminosity. X-ray observations are thus an important component of any multi-wavelength campaign to probe quasar populations. The excellent spatial resolution of the Chandra High Resolution Mirror As- sembly and the effective background rejection of the ACIS instrument make this combination uniquely powerful for quasar surveys. For reference, during a 5 ks observation, the 0.5 -- 8.0 keV background within a 2′′-radius source region is typ- ically ∼0.1 ct. Because ACIS is photon-limited even beyond 100 ks (Alexander et al. 2003), the point-source detection limit scales linearly with exposure time, unlike the √t dependence common in other wavelength bands. In conjunction with sub-arcsec positional accuracy, known optical point sources can be robustly 1 2 Gallagher et al. Hardness ratios versus αox from exploratory Chandra sur- Figure 1. veys for two quasar samples. (a) The Large Bright Quasar Survey broad absorption line quasar sample (Gallagher et al. 2003). (b) The SDSS C iv blueshift sample (Richards et al., in prep.). detected with 3 -- 5 photons. In 5 ks, this corresponds to a 0.5 -- 8.0 keV flux of ∼ 7 × 10−15 erg cm−2 s−1 for a typical quasar X-ray spectrum. A 3 -- 7 ks Chandra exposure, the regime of exploratory observations, is gen- erally insufficient for gathering enough X-rays for spectral analysis of a quasar. However, the strategy of exploratory observations enables the extension of re- sults from spectroscopic observations of individual targets to larger, well-defined samples, and the investigation of connections between X-ray properties and other wavelength regimes. From these datasets, standard X-ray observables are 0.5 -- 8.0 keV flux, hardness ratio,1 and αox.2 We briefly describe the initial results from two exploratory Chandra quasar surveys to illustrate the utility of this observing strategy. Other examples in the literature of successful applications of this approach to quasar studies include surveys of high-z (e.g., Brandt et al. 2002; Vignali et al. 2003), red (Wilkes et al. 2002), and X-ray weak (Risaliti et al. 2003) quasars. 2. X-ray Insights from the LBQS BAL Quasar Chandra Survey We are in the process of performing the largest exploratory survey to date of a well-defined sample of broad absorption line (BAL) quasars drawn from the Large Bright Quasar Survey (LBQS). Since BAL quasars are known to be very faint X-ray sources (e.g., Green & Mathur 1996; Gallagher et al. 1999), ex- 1The hardness ratio is defined to be (h − s)/f , where h=2 -- 8 keV ct, s=0.5 -- 2.0 keV ct, and f =0.5 -- 8.0 keV ct. 2The quantity αox equals 0.384 log(fX/f2500) where fX and f2500 are the flux densities at rest- frame 2 keV and 2500 A, respectively. The Power of Exploratory Chandra Observations 3 Figure 2. Measured X-ray properties from the C4B Chandra survey versus UV properties. (a) αox versus C iv blueshift. (b) Hardness ratio versus ∆(g − i) color. Redder colors are to the right. ploratory observations are the only means of observing sufficient numbers to determine the X-ray properties of the population as a whole. The Chandra data alone are revealing. As seen in Figure 1a, the hardness ratio appears to be anti-correlated with αox. This indicates that the X-ray weakest BAL quasars have the hardest spectra, consistent with the understanding from spectroscopic observations of a handful of objects (e.g., Gallagher et al. 2002) that the X-ray spectra are heavily absorbed. Examining the connection between the X-ray and UV absorption properties of the quasars has also placed observational contraints on quasar disk-wind models (Gallagher et al. 2003). In addition to exploratory Chandra observations, this sample is also being targeted by both SCUBA (PI Priddey) and SIRTF to characterize the submm through hard X-ray spectral energy distributions of BAL quasars as a whole. 3. The Connection Between C iv Blueshift and X-ray Properties High-ionization broad emission lines such as C iv have been known to yield redshifts systematically lower than those measured from Mg ii (e.g., Tytler & Fan 1992); i.e., these C iv lines are blueshifted relative to the systemic velocity. In a study of ∼ 800 SDSS quasars with 1.5 ≤ z ≤ 2.2, Richards et al. (2002) found that the C iv -- Mg ii velocity shifts ranges over ≥ 2000 km s−1. Furthermore, the C iv blueshift (hereafter C4B) was correlated with UV properties, most notably the relative ∆(g − i) color (Richards et al. 2003). That is, the bluest quasars typically exhibited the largest C4Bs. Positing that the C4B might result from the orientation of the accretion disk or the opening angle of the disk wind, we proposed an exploratory Chandra survey to investigate this hypothesis. Six targets, three each from the extreme ends of the C4B distribution, were approved for Cycle 4 observations. While any trends based on six data points need verification, the initial results from this small survey are intrigu- 4 Gallagher et al. ing. Figure 1b shows that hardness appears to increase with αox for the SDSS C4B quasar sample. Though this trend is not statistically significant (Spear- man's rank-order correlation coefficient, rs, is 0.67 for a significance level, prs, of 0.153), the fact that the hardest sources are not X-ray weaker is relevant. This suggests that the hardness of the spectra may not be due to intrinsic absorption in the same way as we see with the BAL quasars. Extending the study to the UV properties, we tested αox versus C4B and hardness ratio versus ∆(g− i) (see Fig- ure 2). Though αox and C4B are consistent with being uncorrelated (rs=−0.77, prs=0.07), the hardness ratio is significantly correlated with ∆(g− i) (rs=−0.99, prs = 3 × 10−4). As shown in Figure 2b, the bluer quasars appear to have softer X-ray spectra, i.e., more negative values of the hardness ratio. This is in line with expectation if UV continuum color is solely related to intrinsic obscuration. However, the lack of connection with αox makes this interpretation uncertain. The connection of UV spectroscopic properties to X-ray emission in these ob- jects implies a physical connection, and more data to investigate this claim are certainly warranted. This experiment also illustrates the value of the SDSS to multiwavelength quasar studies. Given the large number and uniform data quality of the available SDSS quasars, samples can be chosen with precision. Since hardness ratio can vary with z due to absorption and αox is a function of l2500, the luminosity density at rest-frame 2500 A (Vignali et al. 2003), sample tuning significantly reduces potential selection biases. In this C4B quasar survey, the redshifts range from z=1.65 -- 1.89 and l2500 spans only a factor of ∼ 5. The properties of interest, in this case the C iv blueshift and ∆(g − i), are thus more reliably isolated for comparison with the X-ray emission. Acknowledgments. We acknowledge the support of Chandra X-ray center grants GO1 -- 2105X (SCG, WNB) and GO3 -- 4144A (SCG, GTR, WNB). WNB thanks NASA LTSA grant NAG5 -- 13035. References Alexander, D. M., et al. 2003, AJ, 126, 539 Brandt, W. N., et al. 2002, ApJ, 569, L5 Green, P. J., & Mathur, S. 1996, ApJ, 462, 637 Gallagher, S. C., et al. 1999, ApJ, 519, 549 Gallagher, S. C., et al. 2002, ApJ, 567, 37 Gallagher, S. C., et al. 2003, AdvSpRes, in press (astro-ph/0212304) Richards, G. T., et al. 2002, AJ, 124, 1 Richards, G. T., et al. 2003, AJ, 126, 1131 Risaliti, G., et al. 2003, ApJ, 587, 9 Tytler, D., & Fan, X. 1992, ApJS, 79, 1 Vignali, C., et al. 2003, AJ, 125, 2876 Wilkes, B. J., et al. 2002, ApJ, 564, 65 3The significance level ranges from 0.0 -- 1.0 with a small value indicating a significant correlation.
0812.1970
1
0812
2008-12-10T17:17:09
Avoiding the dark energy coincidence problem with a cosmic vector
[ "astro-ph", "gr-qc", "hep-ph" ]
We show that vector theories on cosmological scales are excellent candidates for dark energy. We consider two different examples, both are theories with no dimensional parameters nor potential terms, with natural initial conditions in the early universe and the same number of free parameters as LCDM. The first one exhibits scaling behaviour during radiation and a strong phantom phase today, ending in a "big-freeze" singularity. This model provides the best fit to date for the SNIa Gold dataset. The second theory we consider is standard electromagnetism. We show that a temporal electromagnetic field on cosmological scales generates an effective cosmological constant and that primordial electromagnetic quantum fluctuations produced during electroweak scale inflation could naturally explain, not only the presence of this field, but also the measured value of the dark energy density. The theory is compatible with all the local gravity tests, and is free from classical or quantum instabilities. Thus, not only the true nature of dark energy could be established without resorting to new physics, but also the value of the cosmological constant would find a natural explanation in the context of standard inflationary cosmology.
astro-ph
astro-ph
Avoiding the dark energy coincidence problem with a cosmic vector Jose Beltrán Jiménez and Antonio L. Maroto Departamento de Física Teórica, Universidad Complutense de Madrid, 28040 Madrid, Spain Abstract. We show that vector theories on cosmological scales are excellent candidates for dark energy. We consider two different examples, both are theories with no dimensional parameters nor potential terms, with natural initial conditions in the early universe and the same number of free parameters as L CDM. The first one exhibits scaling behaviour during radiation and a strong phantom phase today, ending in a "big-freeze" singularity. This model provides the best fit to date for the SNIa Gold dataset. The second theory we consider is standard electromagnetism. We show that a temporal electromagnetic field on cosmological scales generates an effective cosmological constant and that primordial electromagnetic quantum fluctuations produced during electroweak scale inflation could naturally explain, not only the presence of this field, but also the measured value of the dark energy density. The theory is compatible with all the local gravity tests, and is free from classical or quantum instabilities. Thus, not only the true nature of dark energy could be established without resorting to new physics, but also the value of the cosmological constant would find a natural explanation in the context of standard inflationary cosmology. Keywords: Dark energy, cosmological vector fields PACS: 95.36.+x, 98.80.-k, 98.80.Es INTRODUCTION The fact that today matter and dark energy have comparable contributions to the energy density, r L ∼ r M ∼ (2 × 10−3 eV)4 in natural units, poses one of the most important problems for models of dark energy. Thus, if dark energy is a cosmological constant, its energy density would remain constant throughout the history of the universe, whereas those of the rest of components (matter and radiation) grow as we go back in time. Then the question arises as to whether it is a coincidence (or not) that they have comparable values today when they have differed by many orders of magnitude in the past. Notice also that if L is a fundamental constant of nature, its scale (around 10−3 eV) is more than 30 orders of magnitude smaller than the natural scale of gravitation, G = M−2 P with MP ∼ 1019 GeV. On the other hand, alternative models in which dark energy is a dynamical component rather than a cosmological constant also require the introduction of unnatural scales in their Lagrangians or initial conditions in order to account for the present phase of accelerated expansion. Such models are usually based on new physics, either in the form of new cosmological fields or modifications of Einstein's gravity [1, 2, 3, 4, 5] and they are generically plagued by additional problems such as classical or quantum instabilities, or inconsistencies with local gravity constraints. Therefore, we would like to find a description for dark energy without dimensional scales (apart from Newton's constant G), with the same number of free parameters as L CDM, with natural initial conditions, with good fits to observations and no consistency problems. In this work we will show that vector theories can do the job. With that purpose, we present two examples of such theories which have been recently proposed [6, 7] (for other vector models see references in those works). SCALING VECTOR DARK ENERGY The action in this case reads [6]: S =Z d4x√−g(cid:18)− R 16p G − 1 4 mn Fmn F ((cid:209) 1 2 − m m A )2 + Rmn A m A n (cid:19) (1) Notice that the theory contains no free parameters, the only dimensional scale being the Newton's constant. The numerical factor in front of the vector kinetic terms can be fixed by the field normalization. Also notice that the "mass" term Rmn Am An can be written as n Am and therefore the theory a combination of derivative terms as (cid:209) contains no potential terms. This action resembles that of Maxwell electromagnetism in the Feynman gauge with a mass term. m Am m An −(cid:209) n An The classical equations of motion derived from the action in (1) are the Einstein's and vector field equations: 1 2 Rgmn Rmn − ✷Am + Rmn A = 8p G(Tmn + T Amn ) = 0 n (2) (3) is the conserved energy-momentum tensor for matter and radiation and T Amn where Tmn is the energy-momentum tensor coming from the vector field. For the simplest isotropic and homogeneous flat cosmologies, we assume that the spatial components of the vector field vanish, so that Am = (A0(t), 0, 0, 0) and that the space-time geometry will be given by: For this metric (3) reads: ds2 = dt2− a2(t)d i jdxidx j, A0 + 3H A0 − 3(cid:2)2H2 + H(cid:3)A0 = 0 (4) (5) Assuming that the universe has gone through radiation and matter phases in which the contribution from dark energy was negligible, we can easily solve this equation in those periods. In that case, the above equation has a growing and a decaying solution: with A±0 constants of integration and a ± = (−3±√33)/6 in the matter era. On the other hand, the (00) component of Einstein's a + + A−0 t A0(t) = A+ 0 t ± = −(1± 1)/4 in the radiation era, and a (6) − a equations reads: H2 = 8p G 3 " (cid:229) a =M,R r a + r A# (7) (cid:209) (cid:209) where the vector energy density is given by: r A = 3 2 H2A2 0 + 3HA0 A0 − 1 2 A2 0 (8) Figure 1: (Left) Evolution of energy densities for the best fit model. Dashed (red) for radiation, dotted (green) for matter and solid (blue) for vector dark energy. We show also for comparison the cosmological constant density in dashed-dotted line. (Right) Evolution of dark energy equation of state for the best fit model. The lower panel shows the 1s confidence interval. Using the growing mode solution from (6), we obtain r A = r A0ak with k = −4 in the radiation era and k = (√33 − 9)/2 ≃ −1.63 in the matter era. Thus, the energy density of the vector field starts scaling as radiation at early times, so that r A/r R = const. However, when the universe enters its matter era, r A starts growing relative to r M eventually overcoming it at some point, in which the dark energy vector field would become the dominant component (see Fig. 1). Notice that since A0 is essentially constant during radiation era, solutions do not depend on the precise initial time at which we specify it. Thus, once the present value of the Hubble parameter H0 and the constant A0 during radiation (which fixes the total matter density W M) are specified, the model is completely determined, i.e. this model contains the same number of parameters as L CDM, which is the minimum number of parameters of a cosmological model with dark energy. As seen from Fig.1 the evolution of the universe ends at a finite time tend where a → aend with aend finite, A0(tend) = MP/(4√p ), r DE → ¥ . This corresponds to a Type III (big-freeze) singularity according to the classification in [8]. and pDE → −¥ We can also calculate the effective equation of state for dark energy as: wDE = pA r A = −3(cid:0) 5 2H2 + 4 3 3 2H2A2 0 + HA0 A0 − 3 H(cid:1) A2 0 + 3HA0 A0 − 1 A2 0 2 A2 0 (9) 2 Again, using the approximate solutions in (6), we obtain: wDE = 1/3 in the radiation era and wDE ≃ −0.457 in the matter era. As shown in Fig. 1, the equation of state can cross the so called phantom divide, so that we can have wDE (z = 0) < −1. In order to confront the predictions of the model with observations of high-redshift supernovae type Ia, we have carried out a c 2 statistical analysis for two supernovae datasets, namely, the Gold set [9], containing 157 points with z < 1.7, and the more recent SNLS data set [10], comprising 115 supernovae but with lower redshifts (z < 1). VCDM Gold L CDM Gold VCDM SNLS L CDM SNLS 0.388+0.023 −0.024 0.309+0.039 −0.037 0.388+0.022 −0.020 0.263+0.038 −0.036 −1 -- −3.53+0.44 −0.48 3.71+0.020 −0.024 −1 -- W M w0 −3.53+0.46 −0.57 A0 (10−4 MP) 3.71+0.022 −0.026 zT 0.265+0.011 −0.012 0.648+0.101 −0.095 0.265+0.010 −0.012 0.776+0.120 −0.108 t0 (H−1 0 ) tend (H−1 0 ) c 2 min 0.926+0.026 −0.023 0.956+0.035 −0.032 0.926+0.022 −0.022 1.000+0.041 −0.037 0.976+0.018 −0.014 -- 0.976+0.015 −0.013 -- 172.9 177.1 115.8 111.0 Table 1: Best fit parameters with 1s intervals for the vector model (VCDM) and the cosmological constant model (L CDM) for the Gold (157 SNe) and SNLS (115 SNe) data sets. w0 denotes the present equation of state of dark energy. A0 is the constant value of the vector field component during radiation. zT is the deceleration-aceleration transition redshift. t0 is the age of the universe in units of the present Hubble time. tend is the duration of the universe in the same units. In Table 1 we show the results for the best fit together with its corresponding 1s intervals for the two data sets. We also show for comparison the results for a standard L CDM model. We see that the vector model (VCDM) fits the data considerably better than L CDM (in more than 2s ) in the Gold set, whereas the situation is reversed in the SNLS set. This is just a reflection of the well-known 2s tension [11] between the two data sets. Compared with L CDM, we see that VCDM favors a younger universe (in H−1 0 units) with larger matter density. In addition, the deceleration-acceleration transition takes place at a lower redshift in the VCDM case. The present value of the equation of state with w0 = −3.53+0.46 −0.57 which clearly excludes the cosmological constant value −1. Future surveys [12] are expected to be able to measure w0 at the few percent level and therefore could discriminate between the two models. We have also compared with other parametrizations for the dark energy equation of state [13]. Since our one-parameter fit has a reduced chi-squared: c 2/d.o. f = 1.108, VCDM provides the best fit to date for the Gold data set. We see that unlike the cosmological constant case, throughout radiation era r DE/r R ∼ 10−6 in our case. Moreover the scale of the vector field A0 = 3.71×10−4 MP in that era is relatively close to the Planck scale and could arise naturally in the early universe without the need of introducing extremely small parameters (for instance in an inflationary epoch), thus avoiding the coincidence problem. In order to study the model stability we have considered the evolution of met- ric and vector field perturbations. Thus, we obtain the dispersion relation and the propagation speed of scalar, vector and tensor modes. For all of them we obtain v = (1− 16p GA2 0)−1/2 which is real throughout the universe evolution, since the value 0 = (16p G)−1 exactly corresponds to that at the final singularity. Therefore the model A2 does not exhibit exponential instabilities. As shown in [14], the fact that the propagation speed is faster than c does not necessarily implies inconsistencies with causality. We have also considered the evolution of scalar perturbations in the vector field generated by scalar metric perturbations during matter and radiation eras, and, again, we do not find exponentially growing modes. If we are interested in extending the applicability range of the model down to so- lar system scales then we should study the corresponding post-Newtonian parameters (PPN). We can see that for the model in (10), the static PPN parameters agree with those of General Relativity [15], i.e. g = b = 1. For the parameters associated to preferred frame effects we get: a 1 = 0 and a 2 = 8p A2 P where A2 ⊙/M2 is the norm of the vector ⊙ ∼ 10−4 (or a 2 < field at the solar system scale. Current limits a 2 < ∼ 10−7 for static vector ∼ 10−5(10−8) M2 fields during solar system formation) then impose a bound A2 P. In or- ⊙ der to determine whether such bounds conflict with the model predictions or not, we should know the predicted value of the field at solar system scales, which in principle does not need to agree with the cosmological value. Indeed, A2 will be determined by ⊙ the mechanism that generated this field in the early universe characterized by its pri- mordial spectrum of perturbations, and the subsequent evolution in the formation of the galaxy and solar system. Another potential difficulty arising generically in vector-tensor models is the presence of negative energy modes for perturbations on sub-Hubble scales. They are known to lead to instabilities at the quantum level, but not necessarily at the classical level as we have shown previously. Work is in progress in order to determine which are the necessary conditions to avoid the presence of such states in this model. < IS THE NATURE OF DARK ENERGY ELECTROMAGNETIC? In the previous section, we have seen that a generic vector field whose action resembles that of electromagnetism with a mass term could be a good candidate for dark energy, but what about standard electromagnetism?. In a very recent work [7], it has been shown using the covariant (Gupta-Bleuler) formalism that it is indeed possible to explain cosmic acceleration from the standard Maxwell's theory. We start by writing the standard electromagnetic action including a gauge-fixing term in the presence of gravity: S = Z d4x√−g(cid:20)− 1 16p G 1 4 R− mn Fmn F + (10) l 2(cid:0)(cid:209) m m A (cid:1)2(cid:21) The gauge-fixing term is required in order to define a consistent quantum theory for the electromagnetic field [16], and we will see that it plays a fundamental role on large scales. Still this action preserves a residual gauge symmetry Am → Am + ¶ m f with ✷f = 0. Electromagnetic equations derived from this action can be written as: mn n F + l m n n A = 0 (11) Notice that since we will be using the covariant Gupta-Bleuler formalism, we do not a priori impose the Lorentz condition. We shall first focus on the simplest case of a homogeneous electromagnetic field Am = (A0(t),~A(t)) in a flat Robertson-Walker background. In this space-time, equations (11) read: A0 + 3H A0 + 3 HA0 = 0 ~A + H ~A = 0 (12) We can solve (12) during the radiation and matter dominated epochs when the Hubble parameter is given by H = p/t with p = 1/2 for radiation and p = 2/3 for matter. In such a case the solutions for (12) are: 0 t + A−0 t−3p A0(t) = A+ ~A(t) = ~A+t1−p +~A− (13) (14) where A±0 and ~A± are constants of integration. Hence, the growing mode of the temporal component does not depend on the epoch being always proportional to the cosmic time t, whereas the growing mode of the spatial component evolves as t1/2 during radiation and as t1/3 during matter, i.e. at late times the temporal component will dominate over the spatial ones. The energy densities of the temporal and spatial components read: H2A2 0 + 3HA0 A0 + r A0 = l (cid:18)9 2 1 ~A)2 2a2 ( ~A = r 1 2 A2 0(cid:19) (15) (16) Notice that we need l > 0 in order to have positive energy density for A0. In fact, it is possible to show that imposing canonical normalization for the corresponding creation and annihilation operators we get l = 1/3 [7]. Besides, when inserting the growing modes of the fields into these expressions we obtain that r A0 = r 0 a−4 A0 and (cid:209) = const. Thus, the field behaves as a cosmological constant throughout the ~A = r 0 m Am , r ~A (cid:209) (cid:209) (cid:209) evolution of the universe since its temporal component gives rise to a constant energy density whereas the energy density corresponding to ~A always decays as radiation. Moreover, this fact prevents the generation of a non-negligible anisotropy which could spoil the highly isotropic CMB radiation. Finally, when the universe is dominated by the electromagnetic field, both the Hubble parameter and A0 become constant (one can straightforwardly check that this is a solution of the complete system of equations) so the energy density is also constant and the electromagnetic field behaves once again as a cosmological constant leading therefore to a future de Sitter universe. As the observed fraction of energy density associated to a cosmological constant today is W L ≃ 0.7, we obtain that the field value today must be A0(t0) ≃ 0.3 MP. The effects of the high electric conductivity s can be introduced using the magneto- hydrodynamical approximation and including the current term Ji = s (¶ 0Ai − ¶ iA0) on the r.h.s. of Maxwell's equations. Notice that because of the universe electric neutrality, conductivity does not affect the evolution of A0(t). The infinite conductivity limit simply eliminates the growing mode of ~A(t) in (14). P, but according to (15), r A0 ∼ H2A2 related to physics at the electroweak scale since r L ∼ (M2 We still need to understand which are the appropriate initial conditions leading to the present value of A0. In order to avoid the cosmic coincidence problem, such initial con- ditions should have been set in a natural way in the early universe. In a very interesting work [17], it was suggested that the present value of the dark energy density could be EW /MP)4, where MEW ∼ 103 GeV. This relation offers a hint on the possible mechanism generating the initial am- plitude of the electromagnetic fluctuations. Indeed, we see that if such amplitude is set by the size of the Hubble horizon at the electroweak era, i.e. A0(tEW )2 ∼ H2 EW , then the correct scale for the dark energy density is obtained. Thus, using the Friedmann equation, we find H2 0 ∼ const., so that r A0 ∼ H4 EW ∼ (M2 A possible implementation of this mechanism can take place during inflation. Notice that the typical scale of the dispersion of quantum field fluctuations on super-Hubble scales generated in an inflationary period is precisely set by the almost constant Hubble parameter during such period HI, i.e. hA2 I [18]. The correct dark energy density can then be naturally obtained if initial conditions for the electromagnetic fluctuations are set during an inflationary epoch at the scale MI ∼ MEW . Despite the fact that the background evolution in the present case is the same as in L CDM, the evolution of metric perturbations could be different, thus offering an observational way of discriminating between the two models. In fact, the evolution of the scalar perturbation F k with respect to the L CDM model gives rise to a possible discriminating contribution to the late-time integrated Sachs-Wolfe effect [19]. The propagation speeds of scalar, vector and tensor perturbations are found to be real and equal to the speed of light, so that the theory is classically stable. We have also checked that the theory does not contain ghosts and it is therefore stable at the quantum level. On the other hand, using the explicit expressions in [15] for the vector-tensor theory of gravity corresponding to the action in (10), it is possible to see that all the parametrized post-Newtonian (PPN) parameters agree with those of General Relativity, i.e. the theory is compatible with all the local gravity constraints for any value of the homogeneous background vector field [20]. EW ∼ M4 EW /MP)4 as commented before. EW /M2 0i ∼ H2 The presence of large scale electric fields generated by inhomogeneities in the A0 field opens also the possibility for the generation of large scale currents which in turn could contribute to the presence of magnetic fields with large coherence scales. This could shed light on the problem of explaining the origin of cosmological magnetic fields. Work is in progress in this direction. CONCLUSIONS We have shown that vector theories offer a simple and accurate description of dark en- ergy in which the coincidence problem could be easily avoided. In our first example, the scaling behaviour during radiation and the natural initial conditions for the vector field offer a neat way around the problem. Moreover, in our second example, the presence of a cosmological electromagnetic field generated during inflation provides a natural explanation for the cosmic acceleration. This result not only offers a solution to the problem of establishing the true nature of dark energy, but also explains the value of the cosmological constant without resorting to new physics. In this scenario the fact that matter and dark energy densities coincide today is just a consequence of inflation taking place at the electroweak scale. Present and forthcoming astrophysical and cosmologi- cal observations will be able to discriminate these proposals from the standard L CDM cosmology. Acknowledgments: This work has been supported by DGICYT (Spain) project num- bers FPA 2004-02602 and FPA 2005-02327, UCM-Santander PR34/07-15875 and by CAM/UCM 910309. J.B. aknowledges support from MEC grant BES-2006-12059. REFERENCES S.M. Carroll, V. Duvvuri, M. Trodden, M.S. Turner, Phys. Rev. D70: 043528, (2004) 1. C. Wetterich, Nucl. Phys. B302, 668 (1988); 2. R.R. Caldwell, R. Dave and P.J. Steinhardt, Phys. Rev. Lett. 80, 1582 (1998) 3. C. Armendariz-Picon, T. Damour and V. Mukhanov, Phys. Lett. B458, 209 (1999) 4. 5. G. Dvali, G. Gabadadze and M. Porrati, Phys. Lett. B485, 208 (2000) 6. 7. 8. J. Beltrán Jiménez and A.L. Maroto, Phys. Rev. D78, 063005 (2008) and arXiv:0807.2528 [astro-ph] J. Beltrán Jiménez and A.L. Maroto, arXiv:0811.0566 [astro-ph] S. Nojiri, S. D. Odintsov and S. Tsujikawa, Phys. Rev. D71 (2005) 063004; M. Bouhmadi-López, P. F. González-Díaz and P. Martín-Moruno, Phys. Lett. B659 (2008) 1 9. A.G. Riess at al. Astrophys.J. 607, 665 (2004) 10. P. Astier et al., Astron. Astrophys. 447: 31-48, (2006). 11. S. Nesseris and L. Perivolaropoulos, JCAP 0702: 025, (2007). 12. R. Trotta and R. Bower, Astron. Geophys. 47: 4:20-4:27, (2006) 13. R. Lazkoz, S. Nesseris, L. Perivolaropoulos, JCAP 0511:010, (2005) 14. E. Babichev, V. Mukhanov and A. Vikman, JHEP 0802, 101 (2008) 15. C. Will, Theory and experiment in gravitational physics, Cambridge University Press, (1993) 16. C. Itzykson and J.B. Zuber, Quantum Field Theory, McGraw-Hill (1980) 17. N. Arkani-Hamed, L. J. Hall, C. F. Kolda and H. Murayama, Phys. Rev. Lett. 85 (2000) 4434 18. A. Linde, Particle physics and inflationary cosmology, Harwood Academic Press (1996) 19. R.G. Crittenden and N. Turok, Phys. Rev. Lett. 76 (1996) 575 20. J. Beltrán Jiménez and A.L. Maroto, arXiv:0811.0784 [astro-ph]
astro-ph/0506459
1
0506
2005-06-20T14:54:40
FADC Pulse Reconstruction Using a Digital Filter for the MAGIC Telescope
[ "astro-ph" ]
Presently, the MAGIC telescope uses a 300 MHz FADC system to sample the transmitted and shaped signals from the captured Cherenkov light of air showers. We describe a method of Digital Filtering of the FADC samples to extract the charge and the arrival time of the signal: Since the pulse shape is dominated by the electronic pulse shaper, a numerical fit can be applied to the FADC samples taking the noise autocorrelation into account. The achievable performance of the digital filter is presented and compared to other signal reconstruction algorithms.
astro-ph
astro-ph
FADC Pulse Reconstruction Using a Digital Filter for the MAGIC Telescope H. Bartko a, M. Gaug b, A. Moralejo c, N. Sidro b for the MAGIC collaboration (a) Max-Planck-Institute for Physics, Munich, Germany (b) Institut de Fisica d Altes Energies, Bellaterra, Spain (c) University and INFN Padova, Italy Presently, the MAGIC telescope uses a 300 MHz FADC system to sample the transmitted and shaped signals from the captured Cherenkov light of air showers. We describe a method of Digital Filtering of the FADC samples to extract the charge and the arrival time of the signal: Since the pulse shape is dominated by the electronic pulse shaper, a numerical fit can be applied to the FADC samples taking the noise autocorrelation into account. The achievable performance of the digital filter is presented and compared to other signal reconstruction algorithms. 1 Introduction The purpose of the MAGIC Telescope [1] is the observation of high energy gamma radiation from celestial objects. When the gamma quanta hit the earth atmosphere they initiate a cascade of photons, electrons and positrons. The latter radiate short flashes of Cherenkov light which can be recorded by a Cherenkov telescope. The FWHM of the pulses is about 2 ns. In order to sample this pulse shape with the 300 MSamples/s FADC system [2], the original pulse is folded with a stretching function leading to a FWHM greater than 6 ns. To increase the dynamic range of the MAGIC FADCs the signals are split into two branches with gains differing by a factor 10. Figure 1a) shows a typical average of identical signals. In order to discriminate the small signals from showers in the energy range below 100 GeV against the light of the night sky (LONS) the highest possible signal to noise ratio, signal reconstruction resolution and a small bias are important. Monte Carlo (MC) based simulations predict different time structures for gamma and hadron induced shower images as well as for images of single muons [7]. An accurate arrival time deter- mination may therefore improve the separation power of gamma events from the background events. Moreover, the timing information may be used in the image cleaning to discriminate between pixels whose signal belongs to the shower and pixels which are dominated by randomly timed background noise. 2 Digital Filter The goal of the digital filtering method [4, 5] is to optimally reconstruct from FADC samples the amplitude and arrival time of a signal whose shape is known. Thereby, the noise contributions to the amplitude and arrival time reconstruction are minimized. 1 a) . ] . u a [ l a n g s i 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0 high gain shape low gain shape MC pulse shape 2 2 4 4 6 6 8 8 10 12 14 10 12 14 time [3.33 ns] time [3.33 ns] ] s n 3 3 . 3 [ e c i l s C D A F 14 12 10 8 6 4 2 0 0 b) 2 4 6 8 10 12 14 FADC slice [3.33 ns] 24 22 20 18 16 14 12 10 8 6 4 2 0 ] 2 ) s t n u o c C D A F ( [ Figure 1: a) Average normalized reconstructed high and low gain pulse shapes and pulse shape imple- mented in the MC simulations [3]. b) Noise autocorrelation matrix B for open camera and averaged over all pixels (galactic telescope pointing). For the digital filtering method, three assumptions have to be made: • The normalized signal shape has to be always constant. • The noise properties must be constant, especially independent of the signal amplitude. • The normalized noise auto-correlation has to be constant. Due to the artificial pulse stretching by about 6 ns on the receiver board all three assumptions are fullfilled to a good approximation. For a more detailed discussion see [8]. Let g(t) be the normalized signal shape, E the signal amplitude and τ the time shift between the physical signal and the predicted signal shape. Then the time dependence of the signal, y(t), is given by y(t) = E · g(t − τ ) + b(t) , where b(t) is the time-dependent noise contribution. For small time shifts τ (usually smaller than one FADC slice width), the time dependence can be linearized. Discrete measurements yi of the signal at times ti (i = 1, ..., n) have the form yi = E · gi − Eτ · gi + O(τ 2) + bi, where g(t) is the time derivative of the signal shape. The correlation of the noise contributions at times ti and tj can be expressed in the noise autocorrelation matrix B: Bij = hbibji − hbiihbji. Figure 2 shows the noise autocorrelation matrix for an open camera. It is dominated by LONS pulses shaped to 6 ns. The signal amplitude E, and the product Eτ of amplitude and time shift, can be esti- mated from the given FADC measurements y = (y1, ..., yn) by minimizing the deviation of the measured FADC slice contents from the known pulse shape with respect to the known noise auto-correlation: χ2(E, Eτ ) = (y − Eg − Eτ g)T B−1(y − Eg − Eτ g) + O(τ 2) (in matrix form). This leads to the following solution for E and Eτ : E = wT amp(trel)y + O(τ 2) with wamp(trel) = ( gT B−1 g)B−1g − (gT B−1 g)B−1 g (gT B−1g)( gT B−1 g) − ( gT B−1g)2 , (1) 2 ) t ( t h g i e w 2.5 2 1.5 1 0.5 0 -0.5 -1 -1.5 -2 -2.5 a) (t) (t) ampw timew pulse shape b) 1 Green Pulses Inner Pixels Blue Pulses Inner Pixels UV Pulses Inner Pixels l ] s n [ n o i t u o s e R e m T i 2c 2c 1T 1T 2T 2T 0T 0T / ndf / ndf [ns]: [ns]: [ns]: [ns]: [ns]: [ns]: 49.92 / 10 49.92 / 10 1.69 0.16 0.16 1.69 0.78 0.78 5.68 5.68 0.21 0.02 0.02 0.21 0.1 0 0 1 1 2 2 4 4 3 3 time [3.33 ns] time [3.33 ns] 5 5 10 3 mean reconstructed signal [ph.el.] 10 2 10 Figure 2: a) Amplitude weights wamp(t0) . . . wamp(t5) for a window size of 6 FADC slices for the pulse shape used in the MC simulations, see text. b) Timing resolution for calibration LED pulses [6]. Eτ = wT time(trel)y + O(τ 2) with wtime(trel) = (gT B −1g)B−1 g − (gT B−1 g)B−1g (gT B−1g)( gT B−1 g) − ( gT B−1g)2 , (2) where trel is the relative phase between g(t) and the FADC clock. Thus E and Eτ are given by a weighted sum of the discrete measurements yi with the weights for the amplitude, wamp(trel), and time shift, wtime(trel), plus O(τ 2). To reduce O(τ 2) the fit can be iterated using g(t1 = t − τ ) and the weights wamp/time(trel + τ ) [4, 8]. Figure 2 a) shows the amplitude and timing weights for the MC pulse shape. The first weight wamp/time(t0) is plotted as a function of trel in the range [−0.5, 0.5[ TADC, the second weight in the range [0.5, 1.5[ TADC and so on. The expected contributions of the noise to the error of the estimated amplitude and timing only depend on the the shape g(t), and the noise auto-correlation B. Analytic expressions can be found in references [4, 8]. 3 Performance and Discussion Figure 2b) shows the measured timing resolution for different calibration LED pulses as a function of the mean reconstructed pulse charge. For signals of 10 photo-electrons the timing resolution is as good as 700 ps, for very large signals a timing resolution of about 200 ps can be achieved. Figure 3 shows the charge and arrival time resolution as a function of the input pulse height for MC simulations (no PMT time spread and no gain fluctuations) assuming an extra-galactic background for different signal extraction algorithms. The digital filter yields the best charge and timing resolution of the studied algorithms [8]. For known constant signal shapes and noise auto-correlations the digital filter yields the best theoretically achievable signal and timing resolution. Due to the pulse shaping of the Cherenkov signals the algorithm can be applied to reconstruct their charge and arrival time, 3 – – – – – – ] . l . i e h p [ ) l a n g s o c e r ( S M R 3 2.5 2 1.5 1 0.5 0 0 0 10 10 10 1 ] s n [ n o i t u o s e r l e m i t 10-1 0 0 10 10 digital filter spline interp. slid. window 20 20 30 30 60 60 input signal [ph.el.] input signal [ph.el.] 40 40 50 50 digital filter spline interp. slid. window 20 20 30 30 60 60 input signal [ph.el.] input signal [ph.el.] 40 40 50 50 Figure 3: Charge and arrival time resolution as a function of the input pulse height for MC simula- tions for the Digital Filter, a cubic spline interpolation (charge= spline integral, time=half maximum position) and a sliding window of 6 FADC slices (charge=samples sum, time = pulse barycenter) [8]. although there are some fluctuations of the pulse shape and noise behavior. The digital filter reduces the noise contribution to the error of the reconstructed signal. Thus it is possible to lower the image cleaning levels and the analysis energy threshold [8]. The timing resolutions is as good as a few hundred ps for large signals. Acknowledgements The authors thank F. Goebel, Th. Schweizer and W. Wittek for discussions and suggestions. References [1] C. Baixeras et al. (MAGIC Collab.), Nucl. Instrum. Meth. A518 (2004) 188. [2] F. Goebel et al. (MAGIC Collab.), in Proceedings of the 28th ICRC, Tokyo, 2003. [3] A. Moralejo et al., MC Simulations for the MAGIC Telescope, In preparation. [4] W. E. Cleland and E. G. Stern, Nucl. Instrum. Meth. A338 (1994) 467. [5] A. Papoulis, Signal analysis, McGraw-Hill, 1977. [6] T. Schweizer et al., IEEE Trans. Nucl. Sci. 49 (2002) 2497. [7] R. Mirzoyan et al., to be published in proceedings of the conference Towards a Network of Atmospheric Cherenkov Detectors VII, 27-29 April 2005 Palaiseau, France. [8] H. Bartko et al., In preparation, to be submitted to Nucl. Inst. Meth. 4
astro-ph/9803280
1
9803
1998-03-24T12:05:01
Jets, plumes and hot spots in the wide-angle tail source 3C130
[ "astro-ph" ]
I present 1.5- and 8.4-GHz observations with all configurations of the NRAO VLA of the wide-angle tail source 3C130. The source has a pair of relatively symmetrical, well-collimated inner jets, one of which terminates in a compact hot spot. Archival ROSAT PSPC data confirm that 3C130's environment is a luminous cluster with little sign of sub-structure in the X-ray-emitting plasma. I compare the source to other wide-angle tail objects and discuss the properties of the class as a whole. None of the currently popular models is entirely satisfactory in accounting for the disruption of the jets in 3C130.
astro-ph
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 16 May 2018 (MN LATEX style file v1.4) Jets, plumes and hot spots in the wide-angle tail source 3C 130 M.J. Hardcastle1⋆ 1 H.H. Wills Physics Laboratory, University of Bristol, Royal Fort, Tyndall Avenue, Bristol BS8 1TL 2 Mullard Radio Astronomy Observatory, Cavendish Laboratory, Madingley Road, Cambridge, CB3 0HE ,2 16 May 2018 ABSTRACT I present 1.5- and 8.4-GHz observations with all configurations of the NRAO VLA of the wide-angle tail source 3C 130. The source has a pair of relatively symmetrical, well- collimated inner jets, one of which terminates in a compact hot spot. Archival ROSAT PSPC data confirm that 3C 130's environment is a luminous cluster with little sign of sub-structure in the X-ray-emitting plasma. I compare the source to other wide-angle tail objects and discuss the properties of the class as a whole. None of the currently popular models is entirely satisfactory in accounting for the disruption of the jets in 3C 130. Key words: radio continuum: galaxies -- galaxies: jets -- galaxies: active -- galaxies: individual: 3C 130 1 INTRODUCTION 3C 130 is a FRI radio source at redshift 0.109 (Spinrad et al. 1985). Its 178-MHz luminosity is 7.6 × 1025 W Hz−1 sr−1, slightly above the nominal FRI-FRII boundary of ∼ 2×1025 W Hz−1 sr−1 (Fanaroff & Riley 1974, hereafter FR). Leahy (1985, 1993) and Jagers and de Grijp (1985) present intermediate-resolution VLA maps of the central regions of the source, while Jagers (1983) has a lower-resolution WSRT image which shows the whole source and its field; the source extends for ∼ 1.5 Mpc. Saripalli et al. (1996) present high- frequency maps made with the Effelsberg 100-m telescope. The host galaxy is classed as a DE2 by Wyndham (1966) and appears to lie in a cluster, although strong galactic red- dening makes optical identification of the cluster members difficult. The Einstein detection of extended X-ray emission (Miley et al. 1983), the nearby aligned sources (Jagers 1983) and the many mJy radio sources in the field at 1.5 GHz make it plausible that the object is the dominant member of a large cluster. Leahy (1985) also attempts to constrain the RM distribution of the source, but notes that it depo- larizes rapidly (particularly in the S lobe) so that few good measurements are available; this could be taken as evidence for a dense magneto-ionic environment for the source (cf. Hydra A, Taylor et al. 1990). 3C 130 is a wide-angle tail (WAT) radio source. The term WAT has been used to describe many different types of object. Here I shall use it to refer to those FRI sources which are associated with central cluster galaxies (e.g. Owen ⋆ E-mail: [email protected] c(cid:13) 0000 RAS & Rudnick 1976) and have luminosities comparable to or ex- ceeding the Fanaroff-Riley break between FRI and FRII. I shall follow Leahy (1993) in using the behaviour of the jets at the base as another defining feature. At high resolution one or two well-collimated jets ['strong-flavour' jets, by the classification of Leahy (1993)] are seen (e.g. O'Donoghue, Owen & Eilek 1990), extending for some tens of kpc before broadening, often at a bright flare point, into the charac- teristic plumes or tails. These jets are very similar to the jets seen in FRII radio galaxies, and quite different from the behaviour of jets in more typical FRIs, where a collimated inner jet, if visible at all, decollimates rapidly (on scales of a few kpc at most) and comparatively smoothly into a bright 'weak-flavour' jet with a large opening angle.† WATs, ac- cording to this definition, never have a weak-flavour jet, but make the transition between strong-flavour jet and diffuse, bent tail in a single step. The requirement that WATs be central cluster galaxies excludes objects (e.g. 3C 171, Blun- dell 1996, Hardcastle et al. 1997a; 3C 305, Leahy 1997) where the 'tails' are likely to be simply ordinary FRII lobes which have been disrupted by unusual host-galactic dynamics. The condition on jet behaviour allows us to exclude objects such as the twin sources in 3C 75 (Owen et al. 1985; Hardcastle 1996) which are associated with a dominant cluster galaxy † There are a few exceptions to this behaviour; 3C 66B (Hardcas- tle et al. 1996) does appear to show an inner 'strong-flavour' jet and a bright knot at the base of the 'weak-flavour' jet. But even here the transition from strong to weak flavours occurs on scales of ∼ 1 kpc. 2 M.J. Hardcastle and sometimes classed as WATs but whose inner jets are similar to those of typical powerful FRIs. Because of the requirements of this definition, wide- angle tail sources make up a small minority of the radio source population. For this reason, the detailed properties of their jets and tails have not been well studied, although a number have been imaged for studies of source dynamics (O'Donoghue et al. 1989). The only objects which have been the subject of detailed study in the radio are 3C 465 (Leahy 1984; Eilek et al. 1984) and 3C 218, Hydra A (Taylor et al. 1990), although M87, Virgo A (e.g. Biretta & Meisenheimer 1993) exhibits some of the properties of a WAT. In this paper I present multi-configuration, multi-frequency VLA observations of a further powerful WAT. Throughout this paper I use a cosmology in which H0 = 50 km s−1 Mpc−1 and q0 = 0. At the distance of 3C 130, one arcsecond is equivalent to a projected length of 2.72 kpc. B1950.0 co-ordinates are used throughout. 2 OBSERVATIONS 3C 130 was observed with the VLA as part of a programme of detailed observations of FRI radio galaxies. Dates and in- tegration times are shown in table 1. 3C 286 and 3C 48 were used as primary flux calibrators; the nearby point sources 0537+531 and 0435+487 were used as phase calibrators, and (where 3C 286 was not observed) 3C 138 was used as a po- larization angle reference. A bandwidth of 50 MHz was used, except at A configuration, where 25 MHz was used to reduce bandwidth smearing. The data were reduced within aips in the standard way. The datasets from each configuration were initially reduced separately, each undergoing several iterations of CLEANing and phase self-calibration. The B, C and D- configuration datasets were then phase-calibrated, using the appropriate baselines, with images made from the higher- resolution datasets, with which they were then merged with- out reweighting. Thus the B-configuration data were phase calibrated with an image made from the A-configuration data and merged with it to form an AB dataset; images made with this at low resolution were used to phase cal- ibrate the C-configuration data and the two were merged to form an ABC dataset, and so on. This process ensures phase consistency in the data while removing the need for a self-calibration of the final merged dataset. Maps were made using the aips task IMAGR, with ta- pering of the uv plane where low-resolution maps were re- quired. The robustness parameter in IMAGR was used to temper the uniform weighting of the uv plane, to improve the signal-to-noise ratio. In all cases the restoring beam was a circularly symmetrical Gaussian, well matched to the Gaus- sian fit to the dirty beam, and the resolution quoted is its FWHM. The total-intensity map at the highest resolution was made with a combination of IMAGR and the maximum- entropy imaging task VTESS; IMAGR was used to clean off the bright point-like components, the residual image was deconvolved with VTESS, and the point-like components subsequently restored. 3 RESULTS 3.1 Overall source structure Fig. 1 shows the large-scale structure of the source. There are several pronounced bends, in spite of the overall straightness of the source. The sudden change in direction at the end of the south tail is particularly noticeable; this feature is similar to several seen in the small sample of O'Donoghue, Owen and Eilek (1990). The source disappears into the noise on these images and is longer than the ∼ 1 Mpc seen here. 3.2 The core The radio core of 3C 130 did not vary over the timescales of the observations either at 8.4 or 1.5 GHz, within the errors imposed by the uncertainty of absolute flux calibration at the VLA. Its flux at 8.4 GHz was 29.0 mJy and at 1.5 GHz 12.4 mJy. The best position for the core is RA 04 48 57.34, DEC +51 59 49.7. 3.3 The jets and hot spots The high-resolution images in Figs 2 and 3 show two very well-collimated jets emerging from the core. The jets are rea- sonably symmetrical. The northern jet in 3C 130 is brighter, noticeably so at bends; over the inner section where both jets are straight (approximately 9 arcsec) the difference in brightness is roughly a factor 1.4. [This symmetry in the brightness of jets is reasonably common among WAT sources, compared to FRII radio galaxies or quasars (e.g. O'Donoghue et al. 1993). The 'archetype' of the class, 3C 465, appears to be unusual in having a very one-sided jet.] The bends in the jets, particularly the northern one, are very striking. The beams may be ballistic, implying some short-timescale wobble of the collimator ('garden-hose' be- haviour), but if this is the case it is surprising that the jets are brighter at bends and that there is no antisymmetry be- tween the jet and counterjet. If they are not ballistic it is equally remarkable that they remain collimated while under- going oscillations of such large amplitude in so short a dis- tance. The northern jet terminates in a hot spot, but there is a long filament which leaves the hot spot to the north, pos- sibly suggesting some continued collimated outflow. At this resolution there is little compact structure at the end of the southern jet; the 'hot spot' seen in the maps of Leahy (1985, 1993) is resolved, with a size of around a second of arc. By contrast, the northern hot spot is only just resolved at the full resolution of the dataset (0.24 arcsec; maps not shown) and its brightest component has a minor axis of ∼ 0.3 arc- sec. This use of the term 'hot spot' is stronger than that of O'Donoghue et al. (1993), who only used it to indicate a brighter, broader region; the hot spot seen here is compara- ble in compactness with those in nearby FRIIs (e.g. Black et al. 1992; Leahy et al. 1997; Hardcastle et al. 1997a) and is su- perposed on a brighter region which corresponds to the 'hot spot' of O'Donoghue et al. The northern jet is resolved at the bends at full resolution, and has a cross-sectional width of up to 0.8 arcsec. The polarization map (Fig. 3) includes a correction for Ricean bias and shows all points with polarized and total intensity greater than three times the respective off-source c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 r.m.s. noise values. The position-angle vectors are perpen- dicular to the observed E-field, and so show the direction of the apparent magnetic field if Faraday rotation is negligi- ble. Although we expect a non-negligible rotation measure (discussed further below, section 3.4), these angles remain the best guess of the magnetic field direction. On this basis, the jets have apparent magnetic field parallel to their length where polarization is detected; the field follows the bends in the northern jet. This is as expected for a strong-flavour jet (e.g. Saikia & Salter 1988). The field in the hot spot is transverse to the jet direction and parallel to the hot spot's direction of extension; this is similar to the field configura- tion in many FRII hot spots (Hardcastle et al. 1997a) but also to that in the termination knots of M87's jet (Owen, Hardee & Cornwell 1989). Further out, the magnetic field is parallel to the plumes, and the degree of polarization is high. This appears to be the behaviour in the best-studied WATs (e.g. Taylor et al. 1990, O'Donoghue et al. 1990, Patnaik et al. 1984; Saikia & Salter 1988, and references therein) but is quite different from the behaviour observed in the weak- flavour jets of normal FRIs, in which the field is transverse to the jet axis, sometimes with a longitudinal sheath (e.g. Hardcastle et al. 1996; Laing 1996; Hardcastle et al. 1997b). 3.4 Depolarization, rotation measure and spectral index Using matched-baseline maps, I confirm earlier findings that the source is rapidly depolarized at low frequencies. The mean depolarization between 1.5 and 8.4 GHz (averaged over the areas with good signal-to-noise in both maps) of the northern plume is 0.2, and that of the southern plume 0.1. It may be noteworthy that the southern lobe, with a weaker jet and no bright compact hot spot, is the more depolarized: this may be an example of a Laing-Garrington effect (Laing 1988; Garrington et al. 1988) in WATs, although Saripalli et al. (1996) suggest that there are substantial variations in the degree of polarization with radio frequency. There is weak evidence that the inner 50 arcsec of both lobes is more depolarized than the outer parts, which would be consis- tent with depolarization by a medium associated with the galaxy or cluster. There are no systematic observations of depolarization in this class of source. The rotation measure (RM) distribution is not con- strained by the rotation of polarization angle between 8.4 and 1.5 GHz. Rotations through all possible angles take place over the source, so there are variations in RM of more than 36 rad m−2 on arcsecond scales. This is consistent with the RM measurements of Leahy (1985). Good maps at a higher frequency are needed to constrain the RM distribu- tion adequately. Saripalli et al. (1996) report measurements suggesting an integrated galactic RM of ∼ 300 rad m−2 in the region of 3C 130. From the fact that the polarization vec- tors are well aligned with one another (and consistent with those in the lower-resolution maps of Saripalli et al.) in the 8.4-GHz maps, and seem to follow bends in the source where these are present, we may guess that the rotation measure towards any point in the source is not much greater than this value, which would produce a 20◦ rotation in polarization position angle at 8.4 GHz. The spectral index of the source steepens rapidly with distance from the core. Fig. 4 shows a map of spectral in- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The WAT source 3C 130 3 dex; the matched baselines of the maps ensure that the steepening is not an effect of undersampling. This spectral behaviour is expected in the standard model in which the plumes flow slowly away from the source [compare the spec- tral index maps of Hydra A by Taylor et al. (1990)]. Note the comparatively flat (α ≈ 0.5) spectral index of the jets and of the material they flow into. The northern hot spot has a spectral index flatter than the material that surrounds it. for I determined spectral ages regions along the (straighter) southern tail, using a minimum energy for the relativistic electron distribution corresponding to γ = 100, an initial electron-energy power-law index of 2 to reflect the hot spot spectral indices of 0.5, no energy contribution from relativistic protons, filling factor unity, and equiparti- tion magnetic fields; I took flux measurements of regions of the plume between 30 and 110 arcsec, measured along the plume, from the radio core. The ageing field used was 0.46 nT, which was the mean of the equipartition fields fitted at various points along the tail; there was little variation in equipartition field strength with distance, so that this field is a good approximation to the correct self-consistent value. The model included the effects of inverse-Compton scattering from the CBR, which at this redshift produces energy loss equivalent to that due to a magnetic field of 0.40 nT. Using a model with effective pitch-angle scattering of electrons (Jaffe & Perola 1973), the plot of age against pro- jected distance along the source was well fitted by a straight line with gradient ∼ 1.2 × 104 km s−1, inferred ages being of the order of 107 years (as found by Jagers & De Grijp 1985). The intercept was non-zero, reflecting the presence of steeper-spectrum material surrounding the jet termination; derived velocities were similar if the intercept was made zero by choosing a steeper initial power-law index (2.53). These inferred ages and velocities in the tails are comparable to those found by spectral age methods in other WAT sources (e.g. Taylor et al. 1990; O'Donoghue et al. 1993) and would imply outflow which is considerably faster than the sound speed in the external medium, given the temperature of the gas around 3C 130 (discussed below); this is perhaps sur- prising in view of the absence of any evidence for post-hot spot shock structures in the tails and of their generally re- laxed appearance. The usual caveats apply to velocities de- termined by spectral-ageing methods, but it should be noted that most factors that can affect the velocity (including a contribution to the energy density from relativistic protons, a particle filling factor less than unity, and significant projec- tion of the radio source) will produce velocities higher than the value given above. Only if the assumptions involved in the spectral ageing analysis are seriously wrong -- for ex- ample, if there is significant in situ particle acceleration in the tails or significant magnetic field inhomogeneity -- can the tail velocity be much lower than this value. Evidence for such processes is discussed in Eilek (1996) and references therein. 3.5 X-ray observations Miley et al. (1983) report on Einstein IPC observations of 3C 130. Serendipitously, the source is also included in the field of a 39.4 ks ROSAT PSPC pointed observation, taken from the public archives, of the X-ray emitting supernova 4 M.J. Hardcastle remnant RX 04591+5147 (Pfeffermann, Aschenbach & Pre- dehl 1991; Reich et al. 1992). Although the X-ray source associated with 3C 130 is 32 arcmin away from the pointing centre of the PSPC, and is thus badly vignetted, the ob- servations have superior signal-to-noise to the Einstein data and show details of the X-ray structure of the source. The cluster is detected at 2200 ± 100 PSPC counts between 0.1 -- 2.4 keV (derived from a circle of 11 arcmin radius about the centre of the X-ray source, using a background annu- lus between 11 and 17.5 arcmin), in spite of the reduced sensitivity of the PSPC at this off-axis distance. Because of the difficulty of measuring the background in the presence of extended emission from the SNR, and because the shad- ows of the ring and one of the radial struts pass close to the source, the derived count rate of ∼ 6×10−2 s−1 is uncertain. A rough correction for vignetting would imply an on-axis count rate of 9 × 10−2 s−1. Using the Post-Reduction Of- fline Software (PROS) within IRAF, I made spectral fits to the data, correcting for the off-axis location of the source. A single Raymond-Smith model provided a good fit (χ2 = 15.8 with 26 degrees of freedom), giving a best-fit temperature kT = 2.9+9 −0.2 × 1022 cm−2 [cf. the value of 0.4 × 1022 cm −2 predicted by in- terpolation from Stark et al. (1992)]. Errors quoted for NH and kT are 1σ for two interesting parameters. Abundances were poorly constrained; 70 per cent solar abundance gave marginally the best fit. With this best-fit model, the 0.1-2.4 keV luminosity of the cluster is 5 × 1037 W, consistent with the luminosity derived, on crude spectral assumptions, from the Einstein data by Miley et al. (1983); the cluster is thus comparable in X-ray luminosity to rich Abell clusters, and the temperature consistent with the temperature-luminosity relation (e.g. David et al. 1993). There is no evidence for a lower temperature in the central regions of the source, and so no evidence that a cooling flow is present; this appears to be normal for the host clusters of WATs (Norman, Burns & Sulkanen 1988; G´omez et al. 1997) although Schindler & Prieto (1997) suggest that a weak cooling flow is present in Abell 2634, the host cluster of 3C 465, and Hydra A inhabits a cooling flow with high mass deposition rates (David et al. 1990). −2 keV; the fitted galactic NH was 0.9+0.5 The best-fit Gaussian to the off-axis point-spread func- tion (PSF) of ROSAT at this distance from the pointing centre has σ ≈ 70 arcsec (Hasinger et al. 1995). For a radio-X-ray comparison I have smoothed the broad-band (0.1-2.4 keV) X-ray image with a Gaussian of this size; this should allow the coma-induced asymmetry of the PSF to be neglected. The X-ray images (Fig. 5) show an extended structure on scales comparable to the length of the radio source (i.e. ∼ 1 Mpc). The cluster gas seems reasonably sym- metrical about the radio source, in contrast to the clumpy structures, with offset radio sources, seen in some lower- luminosity WAT hosts even at lower spatial resolution (e.g. Burns et al. 1994; G´omez et al. 1997). The distortion of the X-ray isophotes to the northeast coincides with, and may be related to, the kink (∼ 150 -- 300 kpc from the nucleus) in the northern tail; there is no structure in the X-ray emission which can be related to the sudden change in direction at the end of the southern radio tail, however. Given the large and asymmetrical PSF, I have not attempted to fit radial profiles to the X-ray data. 4 DISCUSSION Approaches to the source dynamics of WATs in the litera- ture (e.g. Burns 1981; Eilek et al. 1984; O'Donoghue et al. 1993) have concentrated on the large-scale bends seen in the tails. It is instructive to ask a rather different question: why are these sources, with well-collimated strong-flavour jets, compact hot spots, and high radio luminosities, not clas- sical double radio galaxies? In FRII objects of this radio power, radio-emitting plasma is thought to flow back from the hot spots into the 'cocoon' left behind as the hot spot and associated shocks propagate into the external medium, forming the radio lobes (e.g. Scheuer 1974; Williams 1991). In WATs, the jet appears to terminate in a shock in the same way. Norman et al. (1988) argue that strong shocks are necessary to explain the single-step transition between jets and plumes, and observations of compact hot spots in these objects, such as that seen in 3C 130, support this model. However, the situation after the shock is different in the two classes of object. In WATs lobes are not formed. Instead, the hot spot is at the base of the tail; by analogy with the standard model for FRIIs, we may assume that the emitting material in the tail has passed through and been excited in the hot spot, and this is borne out by the spectral index re- sults in 3C 130. The tails may immediately deviate from the axis defined by the jets (e.g. 3C 465) or appear to continue in a straight line (e.g. 3C 130) but in no case does there ap- pear to be lobe emission significantly closer to the core than the hot spot.‡ The fact that there is no cocoon may explain the brightness and two-sidedness of the strong-flavour jets in WATs compared to those in FRIIs; a direct interaction with the (comparatively dense) external medium might be expected to make the beam more dissipative and perhaps to slow the regions of the beam responsible for the emission to only weakly relativistic velocities. This would explain the low values (0.2c) of 'jet velocity' estimated from sidedness by O'Donoghue et al. (1993) compared to the much higher val- ues (0.6 -- 0.7c) estimated from the sidedness of jets in FRII quasars (Bridle et al. 1994; Wardle & Aaron 1997) and their prominence and sidedness in FRII radio galaxies (Hardcastle et al. in prep.). Hardcastle et al. (1997a) proposed a simi- lar explanation for the prominence and two-sidedness of the jets in the peculiar FRII 3C 438. However, in the absence of classical double lobes and the associated discontinuity be- tween radio-emitting plasma and shocked external medium, why are there jet termination shocks in WATs? It is well known that the difference between WATs and classical doubles is the local environment; whereas WATs al- ways lie at the centres of clusters, FRII radio galaxies of com- parable powers tend to avoid them (e.g. Prestage & Peacock 1988). An explanation for the peculiar properties of WATs compared to their classical double counterparts must turn on this environmental difference. A suggestion along these ‡ Whether the bends in the jet seen in 3C 130 are due to ballistic motion or to buffeting by the IGM, it is clear that the position of the jet termination point, however it is formed, must change with time. The hot spot will therefore move about in the base of the plume in a manner similar to that described in the 'dentist's drill' model of Scheuer (1982) for the end points in FRIIs. We do not therefore expect to see the hot spot at a particular place in the tail in all cases. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 lines by Leahy (1984), applied to 3C 465, invoked motion of the galaxy through the cluster, causing it to leave behind a passive wake of radio-emitting material; in this type of model the post-hot-spot material is left behind by the motion of the galaxy and so never forms a lobe. However, the motions of central cluster galaxies are not expected to be large (Eilek 1984; Pinkney et al. 1993 and references therein) and in any case such a model cannot account, without invoking pro- jection effects implausibly often, for the large population of WATs in which one or both tails are more or less aligned with the inner jets (as in 3C 130). Burns et al. (1994) suggest a model in which WATs have an origin in the merger of a cluster with a group or subclus- ter. This is motivated by the X-ray substructure which they find in many WAT host clusters. Large-scale, high-velocity residual motions of gas could then be responsible for the bending of the radio tails, while the merger would provide tidally stripped gas to fuel the AGN. In an extension of this work G´omez et al. (1997) show that the majority of WAT hosts in a larger sample show some X-ray substruc- ture, with an alignment between the direction of the X-ray elongation and the angle that bisects the tails, consistent with such a model. 3C 130, however, is clearly a WAT despite the location of its host at the centre of a smooth, approxi- mately symmetrical distribution of X-ray emitting gas and its (apparently) straight tails. It appears that strong clus- ter inhomogeneity, though it may be necessary for bent tail formation, is not necessary for the existence of a WAT; in particular it does not, on its own, explain the jet shock/hot spot behaviour discussed above. Loken et al. (1995) discuss the physics of a jet propa- gating across the boundary between the interstellar medium of the host galaxy and the hotter, less dense intracluster medium, and suggest that this may be the reason for the disruption of the inner, well-collimated jet at a hot spot. They then postulate velocity shear across the boundary, as described above, to account for jet bending. The structures seen in numerical simulation when the jet simply crosses a contact discontinuity with crosswind do not resemble WATs strongly, however. If the jets are taken to cross a shock front instead (cf. Norman et al. 1988), then the simulations of Lo- ken et al. are more convincing in their resemblance to WATs, but we again face the problem of the smoothness of the large- scale X-ray emission in 3C 130; there is little evidence in this source for the recent cluster merger that Loken et al. invoke to produce such a shock. Neither the sonic radius of a possi- ble cooling flow nor the shock front associated with a puta- tive nuclear or galactic wind are in the appropriate place to produce the internal shocks in WAT jets (Soker & Sarazin 1988; Smith, Kennel & Coroniti 1993). Because of the low resolution of the X-ray data, cluster-merger models cannot be ruled out for 3C 130. Producing such straight plumes in such a model while still causing both jets to disrupt requires a rather special geometry for the merger and/or convenient projection effects, however. If hot spots in WATs represent jet termination shocks, it is perhaps surprising that only a single hot spot is seen in 3C 130 and that there are no clear hot spot candidates in several of the sources of O'Donoghue et al. (1993). It is possible that there are intrinsically similar hot spots but that relativistic beaming effects are affecting their visibility. In 3C 130 the hot spot in the N lobe is approximately ten c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The WAT source 3C 130 5 times brighter than the most comparable feature in the S lobe, which using standard results requires flow or advance velocities greater than 0.3c. More high-resolution observa- tions of these objects are needed to test such a model. 5 CONCLUSIONS A compact hot spot is detected at the base of one plume of the WAT 3C 130, and the jets are shown to have longitudinal magnetic field. The source is thus very like a classical dou- ble in some respects. The data support the model in which WATs are objects whose jets make the transition from super- to sub-sonic velocities in one step, rather than decelerating gradually, by showing a bright sub-kpc structure (compara- ble to those seen in classical double radio sources) associated with the termination of a jet. Archival ROSAT PSPC observations of 3C 130 show it to lie in a luminous cluster with kT ∼ 2.9 keV. There is little sign of substructure in the X-ray, in contrast to many other WATs; this may be related to the nearly straight tails of 3C 130. The lack of strong substructure seems to be in- consistent with recent models for jet disruption in WATs. ACKNOWLEDGEMENTS I am grateful to Julia Riley and Guy Pooley for suggest- ing the original radio observations of this source, and thank Mark Birkinshaw, Julia Riley and Diana Worrall for use- ful comments. I acknowledge a research studentship from the UK Particle Physics and Astronomy Research Coun- cil (PPARC) and support from PPARC grant GR/K98582. The National Radio Astronomy Observatory is operated by Associated Universities Inc., under co-operative agreement with the National Science Foundation. This project made use of Starlink facilities. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is oper- ated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronau- tics and Space Administration. This research has made use of data obtained from the High Energy Astrophysics Science Archive Research Center (HEASARC), provided by NASA's Goddard Space Flight Center. REFERENCES Blundell K.M., 1996, MNRAS, 283, 538 Bridle A.H., Hough D.H., Lonsdale C.J., Burns J.O., Laing R.A., 1994, AJ, 108, 766 Burns J.O., 1981, MNRAS, 195, 523 Burns J.O., Rhee G., Owen F.N., Pinkney J., 1994, ApJ, 423, 94 David L.P., Arnaud K.A., Forman W., Jones C., 1990, ApJ, 356, 32 David L.P., Slyz A., Jones C., Forman W., Vrtilek S.D., 1993, ApJ, 412, 479 Eilek J.A., 1996, in Hardee P.E., Bridle A.H., Zensus J.A., eds, Energy Transport in Radio Galaxies and Quasars, ASP Con- ference Series vol. 100, San Francisco, p. 281 Eilek J.A., Burns J.O., O'Dea C.P., Owen F.N., 1984, ApJ, 278, 37 Fanaroff B.L., Riley J.M., 1974, MNRAS, 167, 31P 6 M.J. Hardcastle Garrington S., Leahy J.P., Conway R.G., Laing R.A., 1988, Nat, 331, 147 G´omez P.L., Pinkney J., Burns J.O., Wang Q., Owen F.N., Voges W., 1997, ApJ, 474, 580 Hasinger G., Boese G., Predehl P., Turner T.J., Yusaf R., George I.M., Rohrbach G., 1995, MPE/OGIP Calibration Memo CAL/ROS/93-015, version 1995 May 08 Hardcastle M.J., 1996, PhD thesis, University of Cambridge Hardcastle M.J., Alexander P., Pooley G.G., Riley J.M., 1996, MNRAS, 278, 273 Hardcastle M.J., Alexander P., Pooley G.G., Riley J.M., 1997a, MNRAS, 288, 859 Hardcastle M.J., Alexander P., Pooley G.G., Riley J.M., 1997b, MNRAS, 288, L1 Jagers W.J., 1983, A&A, 125, 172 Jagers W.J., De Grijp M.H.K., 1985, A&A, 143, 176 Jaffe W.J., Perola G.C., 1973, A&A, 26, 423 Laing R.A., 1988, Nat, 331, 149 Laing R.A., 1996, in Hardee P.E., Bridle A.H., Zensus J.A., eds, Energy Transport in Radio Galaxies and Quasars, ASP Con- ference Series vol. 100, San Francisco, p. 241 Leahy J.P., 1984, MNRAS, 208, 323 Leahy J.P., 1985, PhD thesis, University of Cambridge Leahy J.P., 1993, in Roser H.-J., Meisenheimer K., eds, Jets in Ex- tragalactic Radio Sources, Springer-Verlag, Heidelberg, p. 1 Leahy J.P., 1997, in preparation Loken C., Roettiger K., Burns J.O., 1995, ApJ, 445, 80 Miley G.K., Norman C., Silk J., Fabbiano G., 1983, A&A, 122, 330 Norman M.L., Burns J.O., Sulkanen M.E., 1988, Nat, 335, 146 O'Donoghue A.A., Owen F.N., Eilek J.A., 1990, ApJS, 72, 75 O'Donoghue A.A., Eilek J., Owen F., 1993, ApJ, 408, 428 Owen F.N., Hardee P.E., Cornwell T.J., 1989, ApJ, 340, 698 Owen F.N., O'Dea C.P., Inoue M., Eilek J.A., 1985, ApJ, 294, L85 Patnaik A.R., Banhatti D.G., Subrahmanya C.R., 1984, MNRAS, 211, 775 Pfeffermann E., Aschenbach B., Predehl P., 1991, A&A, 246, L28 Pinkney J., Rhee G., Burns J.O., Hill J.M., Oegerle W., Batuski D., Hintzen P., 1993, ApJ, 416, 36 Prestage R.M., Peacock J.A., 1988, MNRAS, 230, 131 Reich W., Furst E., Arnal E.M., 1992, A&A, 256, 214 Saikia D.J., Salter C.J., 1988, ARA&A, 26, 93 Saripalli L., Mack K.-.H., Klein U., Strom R., Singal A.K., 1996, A&A, 306, 708 Scheuer P.A.G., 1974, MNRAS, 166, 513 Scheuer P.A.G., 1982, in Heeschen, D.S., Wade C.M., eds, Ex- tragalactic Radio Sources, IAU Symposium 97, Reidel, Dor- drecht, p. 163 Schindler S., Prieto M.A., 1997, A&A, 327, 37 Smith S.J., Kennel C.F., Coroniti F.V., 1993, ApJ, 412, 82 Soker N., Sarazin C.L., 1988, ApJ, 327, 66 Spinrad H., Djorgovski S., Marr J., Aguilar L., 1985, PASP, 97, 932 Stark A.A., Gammie C.F., Wilson R.W., Bally J., Linke R.A., Heiles C., Hurwitz M., 1992, ApJS, 79, 77 Taylor G.B., Perley R.A., Inoue M., Kato T., Tabara H., Aizu K., 1990, ApJ, 360, 41 Wardle J.F.C., Aaron S.E., 1996, in Hardee P.E., Bridle A.H., Zensus J.A., eds, Energy Transport in Radio Galaxies and Quasars, ASP Conference Series vol. 100, San Francisco, p. 123 Williams A.G., 1991, in Hughes P.A., ed., Beams and Jets in Astrophysics, Cambridge University Press, Cambridge, p. 342 Wyndham J.D., 1966, ApJ, 144, 459 This paper has been produced using the Royal Astronomical Society/Blackwell Science LATEX style file. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The WAT source 3C 130 7 Table 1. VLA observations of 3C 130 8.4 GHz 1.5 GHz Conf. Date tint Date A B C D 1995/08/06a 1995/11/28 1994/11/10 1995/03/06 (mins) 120 120 55 20 1995/07/23a 1995/11/28 1994/11/10 1995/03/06 tint (mins) 45 30 50 15 a Bandwidth of 25 MHz used. tint denotes the total time spent on source at the specified VLA configuration and frequency. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 8 M.J. Hardcastle Figure 1. 1.5 GHz map of 3C 130 at 10.0-arcsec resolution. Contours at 1.5 × (1, √2, 2, 2√2, . . .) mJy beam−1. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The WAT source 3C 130 9 Figure 2. 8.4-GHz map of 3C 130 at 0.60-arcsec resolution. Linear greyscale; black is 0.4 mJy beam−1. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 10 M.J. Hardcastle Figure 3. 8.4-GHz map of 3C 130 at 0.60-arcsec resolution. Contours at 0.1 × (−2, −1, 1, 2, 4, . . .) mJy beam−1. Vectors show inferred magnetic field direction and their length is proportional to degree of polarization. The inset shows details of polarization in the north jet and hot spot. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The WAT source 3C 130 11 Figure 4. Spectral index between 8.4 and 1.5 GHz of 3C 130 at 2.86-arcsec resolution. Linear greyscale between 0.4 and 1.5; superposed are contours of total intensity at 1.5 GHz at 0.4 × (−2, −1, 1, 2, 4, . . .) mJy beam−1. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 12 M.J. Hardcastle ) 0 5 9 1 B ( N O I T A N I L C E D 52 08 06 04 02 00 51 58 56 54 52 04 49 30 15 RIGHT ASCENSION (B1950) 48 45 00 30 Figure 5. X-ray contours of the 3C 130 cluster, smoothed with a σ = 70-arcsec Gaussian and overlaid on a 10-arcsec resolution greyscale of radio emission at 1.5 GHz (black is 25 mJy beam−1). The lowest contour is at background + 3σ and the contour interval is 2σ. No correction for differential vignetting has been applied; the pointing centre of the PSPC observation was 32 arcmin to the E. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
0710.3355
1
0710
2007-10-17T17:26:47
Lensing Systematics from Space: Modeling PSF effects in the SNAP survey
[ "astro-ph" ]
Anisotropy in the point spread function (PSF) contributes a systematic error to weak lensing measurements. In this study we use a ray tracer that incorporates all the optical elements of the SNAP telescope to estimate this effect. Misalignments in the optics generates PSF anisotropy, which we characterize by its ellipticity. The effect of three time varying effects: thermal drift, guider jitter, and structural vibration on the PSF are estimated for expected parameters of the SNAP telescope. Multiple realizations of a thousand square degree mock survey are then generated to include the systematic error pattern induced by these effects. We quantify their contribution to the power spectrum of the lensing shear. We find that the dominant effect comes from the thermal drift, which peaks at angular wavenumbers l ~ 10^3, but its amplitude is over one order of magnitude smaller than the size of the expected statistical error. While there are significant uncertainties in our modeling, our study indicates that time-varying PSFs will contribute at a smaller level than statistical errors in SNAP's weak lensing measurements.
astro-ph
astro-ph
Lensing Systematics from Space: Modeling PSF effects in the SNAP survey H. F. Stabenau†, B. Jain†, G. Bernstein†, M. Lampton‡ † Dept. of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA ‡ Lawrence Berkeley National Laboratory, Berkeley, CA hstabena, bjain, [email protected], [email protected] Received ; accepted 7 0 0 2 t c O 7 1 ] h p - o r t s a [ 1 v 5 5 3 3 . 0 1 7 0 : v i X r a -- 2 -- ABSTRACT Anisotropy in the point spread function (PSF) contributes a systematic error to weak lensing measurements. In this study we use a ray tracer that incor- porates all the optical elements of the SNAP telescope to estimate this effect. Misalignments in the optics generates PSF anisotropy, which we characterize by its ellipticity. The effect of three time varying effects: thermal drift, guider jitter, and structural vibration on the PSF are estimated for expected parameters of the SNAP telescope. Multiple realizations of a thousand square degree mock survey are then generated to include the systematic error pattern induced by these ef- fects. We quantify their contribution to the power spectrum of the lensing shear. We find that the dominant effect comes from the thermal drift, which peaks at angular wavenumbers l ∼ 103, but its amplitude is over one order of magnitude smaller than the size of the expected statistical error. While there are signifi- cant uncertainties in our modeling, our study indicates that time-varying PSFs will contribute at a smaller level than statistical errors in SNAP's weak lensing measurements. 1. Introduction Current measurements of cosmic shear in the literature have systematic errors at or below the . 10% level, most likely generated by insufficiently corrected optical and atmospheric distortions. Given the accuracy requirements for the shear (between 0.1-1%) for the SNAP survey (Aldering et al. 2004; Albert et al. 2004), it is clear that progress on the optical distortions and shape measurement techniques is essential. The PSF may have spatial and temporal variations and is sampled at only the locations of stars, from where it must be interpolated onto the galaxies. If each exposure had many stars distributed -- 3 -- around the image, then we wouldn't have to worry about time variations in the PSF since there would be enough information in each exposure to subtract the PSF. If the usable star density is too low however, subtracting a single static PSF from every exposure will be inadequate. Recent work shows that the many exposures used in the planned SNAP survey could be used to make this correction very precisely (Jarvis & Jain 2004). In this paper, we estimate the shear power spectra for a fully dynamic treatment of the observatory with thermal and mechanical PSF disturbances that continue throughout the one year survey. To model a first-order PSF correction, we use an imperfect static PSF that is held fixed throughout the same one year survey. A further improvement not pursued here would be to update the PSF model throughout the survey using daily field star images. Our shear power spectra should therefore be regarded as upper limits to the corrected shear spectra, given the model parameters we have assumed. An estimate of the limiting systematics for ground- and space-based WL data requires some knowledge of the amplitude and variability of the PSF. We are working with other members of the SNAP Weak Lensing Working Group (WLWG), and the SNAP optical-mechanical engineers, to model the temporal behavior of its PSF. In this section we describe our work on systematic errors: calculation of the expected instrumental distortion for SNAP by modeling the dominant sources of PSF anisotropy. PSF anisotropy and time variation have been the dominant systematic errors in most lensing measurements (Refregier 2003). This is one reason why a space instrument is expected to be superior for lensing. The expected variations, e.g. thermal effects, should be very slow, given the absence of gravity and the nearly constant thermal environment. It is therefore important to quantify the time rate of change of the PSF for a nominal SNAP mission to determine the degree to which the recovered shear maps can be corrected for instrumental effects. An initial assessment of the SNAP PSF time variations was presented by Sholl et al. (2005). -- 4 -- 2. PSF of the SNAP telescope We have calculated the PSF ellipticity on a grid over the field of view of the instrument, subject to specified optical misalignment parameters. The first step is to use a ray-tracing telescope simulator for SNAP to generate monochromatic PSF images at one micron wavelength as a function of position of the field of view and telescope misalignment. The ray tracer filled the aperture with a grid of rays from a distant point source, propagating them through the optical system to evaluate the pupil's complex transmission, and then Fourier transforming and squaring to get the image irradiance. Finally, we determined the second moments of the image and from them its ellipticity components. The resulting pattern of PSF anisotropy on the focal plane for a sample misalignment is shown in Figure 1. The length and orientation of each whisker represents the magnitude and orientation of the ellipticity at that position. The typical ellipticity shown is ∼ 0.2%. We define the ellipticity in terms of the second moments of the diffracted light, e1 = e2 = Ixx − Iyy Ixx + Iyy 2Ixy Ixx + Iyy . The second moments Iab for a, b ∈ {x, y} are defined as Iab = Pj Ijajbj Pj Ij , where Ij are the intensity values of the image pixels, and xj, yj are the coordinates of a pixel on the focal plane. The second moments Ixx etc. were computed on an unweighted basis for this study. Applying a weighting kernel appropriate for individual galaxy sizes would slightly reduce our moments but not significantly influence our conclusions. Whiskers are plotted at an angle of with respect to the x-axis, so that -- 5 -- θ = 1 2 arctan e2 e1 • pure positive e1 corresponds to a horizontal whisker ( -- ) • pure negative e1 corresponds to a vertical whisker () • pure positive e2 corresponds to a 45 deg whisker (/) • pure negative e2 corresponds to a 45 deg whisker (\) We next consider the process of generating a survey with SNAP, including effects that cause the PSF ellipticity pattern to change over the length of the survey. This could happen if there were some time-varying instrumental effects, which would translate into a systematically changing PSF across the field of the survey. We modeled a step and repoint survey by taking an exposure using the chips for one filter band, regenerating the PSF, moving the telescope by one CCD chip width, and repeating the process, resulting in a series of strips that we then put together to form a 32 × 32 square degree mock survey. Multiple realizations of this survey were used to compute the power spectrum. We considered three time-varying misalignment effects when generating the survey, categorized by the timescale of their variation: • Thermal drift, which has a timescale of days to weeks; • Telescope structural vibration, which has a timescale of an hour; and • Telescope guide jitter, which is different for every exposure. -- 6 -- Of these three effects, the thermal drift effect dominates the time-varying component of the PSF, while the effect of telescope guide jitter turned out to have almost no measurable contribution. The thermal contributions to the PSF include a daily heat pulse from daily attitude maneuvers for telemetry, plus long term drifts from slow changes in the solar longitude and from control system drift. We ignore the daily environmental component because the thermal response is dominated by the thermal control loop on this time scale. We attempted to measure the slower thermal drift contribution to the PSF power spectrum model by modeling a drift in the control system of the sensors that measure the temperature of the struts that hold up the secondary mirror, which undergo thermal expansion and contraction. If the struts have varying lengths, the secondary mirror will be defocussed, decentered, and tilted. The SNAP focal plane includes dedicated star tracker image sensors to determine the instantaneous attitude (Secroun et al. 2002). By analyzing the positions of stars within an image and comparing to catalogs, the pointing of the telescope is established and corrected via a feedback mechanism with telescope orientation controls. However Poisson noise makes the centroids of stars in the image appear to shift around. The attitude control system responds to this centroid jitter according to its control bandwidth, which cannot be zero. We modelled the jitter by assuming a continuing stream of star centroid information from the star guiders, furnished at a rate of 10 frames/second, with an individual centroid RMS noise level of 1 milliarcsecond per axis, as described by Secroun et al. (2002). Because these successive centroid determinations are based on independent photon arrivals, the deviations will be statistically independent and circularly symmetric around the true long term mean. We assume that the spacecraft attitude control bandwidth is of the order of 0.1 Hz, so that ∼ 100 frames are averaged in establishing each attitude, and the RMS jitter in each axis will be reduced tenfold from the individual frame centroids, to ∼ 0.1 milliarcsecond. The guider's effect on the PSF turned out to be much smaller than both -- 7 -- Source Amplitude Angular scale thermal decentering along x, y directions 0.8µm 7 milliarcsec thermal defocus along z direction 0.5µm 4.6 milliarcsec vibrational tilt guide jitter 1.6µm 16 milliarcsec 0.01µm 0.1 milliarcsec. Table 1: Comparison of the RMS widths of the Gaussian distributions from which the misalignment parameters are chosen during the simulated survey. the vibrational and thermal effects. Vibrational modes are excited in the structure of the telescope by the vibration of four flywheels that store angular momentum used to rotate the satellite in space. This causes the secondary mirror to vibrate with an amplitude which changes as the flywheel goes in and out of resonance with the telescope structure. We model the vibration effect as the sum of two tilted orientations of the secondary mirror, since the structural vibration modes of the SNAP observatory have frequencies above 15 Hz, and WL exposures have durations of greater than 300 seconds, the number of vibration cycles in each exposure is large, > 4500, allowing us to average the PSF over many cycles. First, we choose an amplitude for the vibration from a Gaussian distribution. Next, we generate two exposures each with the secondary mirror tilted by plus and minus this amplitude. Finally, we average the images from the two orientations. The result is an exposure affected by our model vibrational PSF. Spacecraft systems engineers provided projections of the magnitude of the thermal and vibrational effects (Ericsson & Stoneking 2001). In order to simulate surveys affected by time-varying misalignments, during each model survey new misalignment parameters for the secondary were chosen from Gaussian distributions every two weeks during the model surveys. The widths of the Gaussian distributions used for the misalignments of the secondary are shown in Table 1. -- 8 -- At intermediate times the misalignments are linearly interpolated between the old and new values. One such realization of the set of model surveys can be seen in Figure 2. The RMS shear in that figure is about 0.1%. 3. Shear Power Spectrum The systematic error contribution to the lensing shear power spectrum can be obtained by Fourier transforming of the gridded PSF ellipticity maps described above. The power spectrum due to the thermal drift is shown in Figure 4, along with the power spectrum due to vibrations, which has a much smaller amplitude. The power spectrum due to jitter is smaller still, and is below the scale plotted. For the thermal drift, the full contribution to the power spectrum, and the smaller contribution due the residual shear after the static pattern is subtracted are shown. This subtraction is important; for real data it is expected that the static PSF can be measured accurately by interpolating from stars in the image (Jarvis & Jain 2004). In order to compare the effects of PSF anisotropy with the lensing signal, we generated some simulated shear fields via ray-tracing through N-body simulation boxes (Heitmann et al. 2004). The simulations were performed in a ΛCDM cosmology with a comoving box size of 500 h−1Mpc, and the sources were set at a redshift of z = 2. Figure 3 shows a whisker plot of one of the realizations; the RMS shear in the field shown is about 2%. Figure 4 shows that the power spectrum of the time-varying residual of the PSF is typically smaller than the lensing signal by three to four orders of magnitude. It is also much smaller than the statistical errors expected for a thousand square degree survey, also shown in the plot. These include sample variance and the shot noise contribution of galaxies with an RMS ellipticity of σǫ = 0.4 (both components combined) and number density ng = 100 per square arcminute. Thus for the standard parameters of the telescope, -- 9 -- we have shown that the power spectrum of the residual error due to PSF anisotropy makes a negligible contribution to the lensing power spectrum. 4. Discussion We have estimated the effect of three time varying components of PSF anisotropy to the weak lensing power spectrum for the optical design of the SNAP telescope. The time variation of the three effects considered, thermal drift, guider jitter and structural vibrations, leads to spatial patterns in the shear maps inferred from the measured ellipticities of galaxies. We find that for the current best estimates of the amplitude of these three effects, the contribution to the lensing power spectrum is much smaller than the expected statistical errors (after subtracting the static component of the PSF pattern). The systematic error contribution due to PSF anisotropy is therefore negligible, excluding catastrophic failures, revised estimates of the effects we have modeled, or other effects not included in this study. It would further be of interest to compute the bispectrum of the PSF pattern and compare it to the signal for different triangle configurations. The bispectrum carries important cosmological information, so it is important to test how it will be affected by systematic errors. Preliminary results for special configurations suggest that the contribution to the bispectrum is also much smaller than the expected signal. We will further include worst-case sources of misalignment and catastrophic events such as an extended power outage. Finally, we note that even in the presence of large systematic errors, the estimate of cosmological parameters may be compromised only to a limited extent, as the lensing signal for multiple auto and cross-spectra will in general scale differently with these parameters (Huterer et al. 2006). -- 10 -- We thank Katrin Heitmann, Salman Habib, and Derek Dolney for help with N-body simulation data. We acknowledge useful suggestions from Mike Sholl, Michael Levi, and Saul Perlmutter. -- 11 -- Fig. 1. -- Anisotropy in the point spread function (PSF) in the SNAP focal plane. Possible misalignments in the telescope optics have been modeled by ray tracing. In this figure we have modeled a large misalignment: the secondary mirror is off-center by 10µm. The PSF at each position is obtained by computing the second moment of the intensity. The RMS size of the whiskers shown represents a 0.2% ellipticity. -- 12 -- Fig. 2. -- PSF anisotropy generated by thermal drift in the SNAP telescope. By choosing mis- alignment parameters from a Gaussian distribution every two weeks, we obtain a spatially varying PSF pattern with a characteristic length scale. The whiskers in this plot have an RMS ellipticity at the ∼ 0.1% level. Compare with Figure 3, which contains simulated weak lensing whiskers, and Figure 4, which plots the power spectra. Subtracting off the static pattern produces residuals that are smaller by about an order of magnitude, as shown in Figure 4 for the power spectrum. -- 13 -- Fig. 3. -- Ray-tracing through N-body simulation data provides a way to estimate the importance of the time-varying SNAP PSF. The power spectrum of the ensemble of simulations is the green curve in Figure 4. -- 14 -- Fig. 4. -- Power spectra of the lensing signal and expected PSF anisotropy contamination for the SNAP telescope. Thermal effects and vibration in the structure of the telescope lead to time variation in the PSF anisotropy. This residual power would lead to an additive systematic error in the lensing measurement. Subtracting the static pattern (see text) potentially provides a reduction in PSF power of as much as an order of magnitude. The red curve (squares) is the time-varying power due to thermal effects, and has not had the static PSF pattern subtracted off, whereas the blue curve (triangles) is the residual power after the time invariant pattern is subtracted (as it is expected to be measured accurately). The statistical errors for a 1000 square degree survey are shown by the green squares, along with the power spectrum amplitude of the shear signal from simulations (circles). The statistical error is projected to be larger by > 1 orders of magnitude than the time-varying residual PSF signal. The much smaller residual power from structural vibrations is about two orders of magnitude below the bottom of the plot. -- 15 -- REFERENCES Aldering, G., et al.: PASP, submitted, astro-ph/0405232 (2004) Ericsson, A., Stoneking, E.: NASA IMDC Report for SNAP, June 25 -- 28, Chart 12 (2001) Heitmann, K., Ricker, P. M., Warren, M. S., Habib, S.: Astrophys. J. Suppl. 160, 28 (2005) Huterer, D., Takada, M., Bernstein, G., Jain, B.: MNRAS 366, 101 (2006) Jarvis, M., Jain, B., ApJ, submitted, astro-ph/0412234 (2004) Refregier, A.: ARA&A 41, 645 -- 68 (2003) J. Albert et al. [SNAP Collaboration]: Astropart. Phys. 20, 377 (2004) Secroun, A. et al.: Experimental Astronomy 12 (2) 69 -- 85 (2001). Sholl, M. et al.: Proc. SPIE v.5899 27 -- 38 (2005). This manuscript was prepared with the AAS LATEX macros v5.2.
astro-ph/9803241
1
9803
1998-03-20T12:35:35
Possible Production of High-Energy Gamma Rays from Proton Acceleration in the Extragalactic Radio Source Markarian 501
[ "astro-ph" ]
The active galaxy Markarian 501 was discovered with air-Cerenkov telescopes at photon energies of 10 tera-electron volts. Such high energies may indicate that the gamma rays from Markarian 501 are due to the acceleration of protons rather than electrons. Furthermore, the observed absence of gamma ray attenuation due to electron-positron pair production in collisions with cosmic infrared photons implies a limit of 2 to 4 nanowatt per squaremeter per steradian for the energy flux of an extragalactic infrared radiation background at a wavelength of 25 micrometers. This limit provides important clues on the epoch of galaxy formation.
astro-ph
astro-ph
Possible Production of High-Energy Gamma Rays from Proton Acceleration in the Extragalactic Radio Source Markarian 501 K. Mannheim Universitats-Sternwarte, Geismarlandstrasse 11, Gottingen D-37083, Germany ([email protected]) SCIENCE 279, 684 (1998) 8 9 9 1 r a M 0 2 1 v 1 4 2 3 0 8 9 / h p - o r t s a : v i X r a Karl Mannheim 2 ABSTRACT The active galaxy Markarian 501 was discovered with air-Cerenkov telescopes at photon energies of 10 tera-electron volts. Such high energies may in- dicate that the γ rays from Markarian 501 are due to the acceleration of protons rather than electrons. Furthermore, the observed absence of γ ray attenuation due to electron-positron pair production in collisions with cosmic infrared photons implies a limit of 2 to 4 nanowatt per squaremeter per stera- dian for the energy flux of an extragalactic infrared radiation background at a wavelength of 25 micrometers. This limit provides important clues on the epoch of galaxy formation. Karl Mannheim 3 Gamma rays (γ rays) from cosmic sources impinging on Earth's atmo- sphere initiate electromagnetic showers in which the energy of the primary γ ray is imparted among secondary electron-positron pairs. The blue Cerenkov light emitted by the pairs in the atmosphere can be detected from the ground with optical telescopes triggering on the short (∼ 1 ns) optical pulses. The technique has advanced considerably in recent years (1), and some surpris- ing discoveries have been made. Among them is the detection of the blazar Markarian 501 (Mrk 501) at energies above 10 TeV (1 TeV = 1012 eV) (2). Blazars are remote but very powerful sources characterized by their vari- able polarized synchrotron emission. They are associated with radio jets (bipolar outflows) emerging from giant elliptical galaxies seen at small an- gles with the line of sight. Mrk 501 is ∼ 3 × 108 light years from Earth but nevertheless produces a tera-electron volt γ ray flux during outbursts that is many times stronger than that of the Crab nebula, a supernova remnant inside our Milky Way at a distance of only 6 × 103 light years. The radiation mechanism responsible for the γ rays could be either inverse-Compton scat- tering of low-energy photons by accelerated electrons (3) or pion production by accelerated protons. In the latter case, the sources could be among the long-sought sources of cosmic rays, that is the isotropic flux of relativistic particles with differential number density (N) spectrum dN/dE ∝ E −2.7 (for energies E < 103 TeV) mainly consisting of protons and ions (4). Particle acceleration in astrophysics is typically observed to be associated with (collisionless) shock waves when a supersonic flow of magnetized ma- terial hits a surrounding medium. Examples of shock waves are shell-type supernova remnants (explosion of a massive star), plerions (pulsar wind), γ ray bursts (relativistic ejecta from the collapse of a compact stellar object), or the jets ejected from active galactic nuclei (collimated relativistic wind from the accretion disk around a supermassive black hole). Karl Mannheim 4 In the theoretical picture of shock acceleration, relativistic particles (pro- tons, ions, electrons) scatter elastically off turbulent fluctuations in the mag- netic field on both sides of the shock and thereby gain energy because of the convergence of the scattering centers (approaching walls). The acceleration time scale for the process can be written as tacc = ξrgc/u2 where u denotes the velocity of the shock wave (c is the speed of light) and rg ∝ E/B denotes the radius of gyration of a particle with energy E in a magnetic field of strength B. The effects of shock obliquity, turbulence spectrum, and other unknowns are conveniently hidden in an empirical factor ξ ≥ 1. The most rapid (gyro- time scale) particle acceleration for relativistic shocks corresponds to ξ = 1 (5). Balancing the acceleration time scale with the energy loss time scale due to synchrotron radiation tsyn ∝ B−2E −1 one obtains the maximum en- ergy of the electrons Emax = 10 (ξ/10)−0.5(B/3µG)−0.5(u/108 cm s−1) TeV (6). The observed 10 TeV γ rays from the Crab nebula (7) and the observed synchrotron x-rays in shell-type supernova remnants (8) (corresponding to 10 TeV electrons) require ξ ∼ 1 to 10. Because protons lose less energy, they can reach larger Emax's than electrons and give rise to γ ray emission even above ∼ 10 TeV via pion production and subsequent pion decay. Although shock acceleration theory predicts that most of the cosmic rays are acceler- ated in supernova remnants (4), no definitive γ ray signature has yet been discovered. It is commonly argued that the assumption of electron acceleration also suffices to explain the γ rays from blazar jets such as Mrk 501 (9). Estimates of the magnetic field strength in the γ ray emitting part of the jet in Mrk 501 then yield values in the range B ∼ 0.04 to 0.7 G. This magnetic field is much stronger than the one in supernova remnants and the associated stronger cooling of the relativistic electrons due to synchrotron energy losses reduces Emax accordingly. The effect is almost compensated by the high shock wave Karl Mannheim 5 velocities in extragalactic radio sources which speed up the acceleration rate. Using radio interferometry, shock wave velocities close to the speed of light have been inferred corresponding to typical bulk Lorentz factors in the range Γjet = (1 − β 2)−0.5 ∼ 2 to 10 (β = u/c) with a few cases of still higher values (10). Due to the alignment of the jet axis and the line-of-sight in Mrk 501 superluminal motion has not been observed. With u = c, ξ = 10, and tak- ing into account equal synchrotron and inverse-Compton losses, one obtains Emax ∼ 4 Γjet(B/G)−0.5 TeV from the balance between acceleration gains and energy losses. The additional factor Γjet takes into account the boost in energy due to the relativistic bulk motion. Therefore, electron maximum energies of ∼ 10 TeV as required for Mrk 501 (at least 5-8 TeV are required for the similar blazar Mrk 421 (11)) are formally allowed, but one certainly has to push the theory to its limits and this raises a number of concerns: (i) the multi-TeV spectrum should show considerable curvature due to the (so- called Klein-Nishina) decrease of the scattering cross section when the energy of the scattered photon approaches Emax and due to the onset of electron- positron pair production (the observed multi-TeV spectrum is consistent with a smooth power law), (ii) the ratio between the γ ray and synchrotron (si- multaneous) luminosities depends sensitively on the jet Lorentz factor Γjet and therefore requires fine-tuning (both nearest bright blazars, Mrk 421 and Mrk 501, show a similar γ-to-x-ray luminosity ratio), (iii) the magnetic field pressure turns out to be much lower than the relativistic electron pressure in the electron acceleration models which seems inconsistent with the shock acceleration mechanism (the turbulent magnetic field is responsible for push- ing the electrons back and forth across the shock), larger values of B are also expected from the adiabatic expansion of a magnetically collimated jet (the observed B-field at the tips of the jet in Cygnus A is consistent with adiabatic expansion (12)), and (iv) larger values of Γjet could ameliorate the problem that Emax ∼ 10 TeV, however, one would run into a serious problem with unification models of active galaxies if Γjet > 10 would be the rule rather Karl Mannheim 6 than the exception (13) (number of required host galaxies would exceed the number of known radio galaxies). A natural solution to the problem is to assume that the 10 TeV γ rays are due to pion production from accelerated protons. The balance equa- tion between energy gains and losses for protons yields maximum energies of ∼ 106 TeV and short variability time scales tvar ≥ 105(ξ/Γjet)(B/G)−1 s in Mrk 501 (14). The relativistic proton energy loss is dominated by photo- production of pions in collisions with low-energy synchrotron photons (origi- nating from accelerated electrons). Collisions of the accelerated protons with matter are negligible due to the low matter density in relativistic jets (unless a high-density target moves across the jet (15)). The γ rays from the decay of the neutral pion (far above the observed range of energies) are subject to pair creation in further collisions with the low-energy synchrotron photons (γ + γ → e+ + e−). This initiates an electromagnetic cascade shifting the average photon energy to the TeV range and below. A model based on the combined acceleration of protons (γ rays) and electrons (radio-to-x-rays) -- coined the proton blazar model (16) -- was fitted to published data of Mrk 501 in order to obtain a prediction of its multi-TeV spectrum (17). Data from 1995 and earlier was available for the analysis and covered the radio-to-x-ray wavelength range including a flux limit above 100 MeV and an integral flux above 300 GeV (for details see references in (17)). The published flux values showed considerable variability in the optical-to-x-ray range and the model spectrum was therefore fitted to match the time-averaged spectrum (from the fit one obtains B = 37 G and Γjet = 10). The predicted multi-TeV spec- trum is shown in Fig.1 and compared with the data obtained from recent (1996, 1997) air-Cerenkov observations. The fairly robust spectral slope of the model spectrum fits the observations very well, whereas the absolute flux normalization is somewhat too low. Considering that the sub-TeV (350 GeV) flux has been reported to increase from 1995 to 1997 (9), the agreement is Karl Mannheim 7 actually rather impressive if one scales up the model spectrum accordingly. Contemporaneous multi-wavelength observations of blazars such as Mrk 501 will be important to discriminate between electron-based and proton-based models for the γ ray emission from them. Generally, electron-based models require larger values of Γjet and lower values of B than proton-based models to obtain high-energy γ rays. There is a further hint that proton acceleration might be important. Un- resolved blazars are the most probable source population to produce the observed extragalactic γ ray background between 10 MeV and 10 GeV (18). The energy density of this γ ray background is ≃ 4 × 10−6 eV cm−3. A sim- ilar value is found for an extragalactic flux of protons (with dN/dE ∝ E −2 differential spectrum between 109 eV and 1020 eV) providing all of the ob- served cosmic rays with energies above 1018.5 eV (the so-called "ankle" above which the very high energy differential cosmic ray spectrum ∝ E −3 flattens) (19). On the assumption that the γ rays are from proton (p) acceleration, the comparable energy densities result from simple decay kinematics: Photo- production of neutral pions (π0) (p + γ → π◦ + p) and their subsequent decay gives rise to γ rays (subject to electromagnetic cascading) which carry ∼ 1/5 of the proton energy. Charged pions (p + γ → π+ + n) are produced with approximately the same rate and give rise to neutrons (n) which carry ∼ 4/5 of the accelerated proton energy. Since the neutrons do not scatter off the magnetic field fluctuations responsible for the acceleration and storage of the charged particles in the blazar jet, they must escape the accelerator (energy losses are small and modify the neutron spectrum only at the very highest energies). Time dilation allows the most energetic neutrons to leave the host galaxy freely before β-decay occurs (for protons there would be adiabatic losses due to the magnetic field in the host galaxy). Hence the neutron (and after β-decay proton) luminosity is equal to the γ ray luminosity within fac- tors of order unity. An extragalactic origin of the highest energy cosmic rays Karl Mannheim 8 is indeed suggested by the absence of an enhancement of the cosmic ray flux toward the Galactic disk (20) and by the change in chemical composition from heavy (protons and ions) to light (protons) above 1018.5 eV (21). The ultimate challenge to the hypothesis is the measurement of the high-energy neutrino (ν) flux associated with the charged pion decay (π± → e± + 3ν). The energy density in these multi-TeV neutrinos would be of the same order of magnitude as that in extragalactic cosmic rays and γ rays. Their mea- surement therefore constitutes an experimentum crucis within reach for the planned cubic kilometer underwater(ice) detectors (22). Although γ rays are known from laboratory experiments for their pene- trating power, propagation over intergalactic distances is not without hurdles. A diffuse isotropic infrared background (DIRB) was produced when the first galaxies formed. Massive stars in early galaxies produced large amounts of dust in their winds reprocessing the visual and ultraviolet light from the stars into infrared light. By colliding with these ample infrared photons, γ ray photons can disappear and turn into electron-positron pairs (23-25). The most numerous infrared photons above threshold for pair production with 10 TeV γ rays have wavelengths ∼ 25 µm. The mean free path (λγγ) for pair creation at multi-TeV energies is of the order of the distance (d) of Mrk 501. The exact value depends on the DIRB which is difficult to mea- sure directly owing to the presence of zodiacal light and galactic cirrus clouds. One can use the observed power law spectrum (2) to put a limit on the maximum allowed pair attenuation assuming that the observed power law is the unattenuated spectrum emitted by the source (consistent with the proton-based model). In general, only very contrived intrinsic spectra would look like a smooth power law after the quasi-exponential attenuation. The maximum allowed deviation from the power law (1 − exp[−d/λγγ]) is taken to be the size of the statistical error bar at 10 TeV yielding an optical depth Karl Mannheim 9 τγγ = d/λγγ < 0.7. This limit can be relaxed by a factor not larger than ∼ 2 admitting for weakly absorbed spectra which still approximate a power law (see the dashed line in Fig.1). There is some dependence of the attenuation on the shape of the DIRB spectrum. Useful models for the spectral shape can be found in (23-25) and yield a similar limit for the 25 µm DIRB normaliza- tion νIν(25 µm) < 2 to 4 nW m−2 sr−1. The absence of γ ray attenuation in Mrk 501 is consistent with no contribution to the DIRB other than from the optically selected galaxies for which one expects ∼ 10% of their optical emis- sion to be reprocessed by warm dust yielding νIν(25µm) ∼ 1 nW m−2 sr−1 (26), but would also allow a DIRB stronger by a factor 2 to 4. A DIRB of at least ∼ 3 nW m−2 sr−1 is suggested by faint infrared galaxy counts and indicates contributions from dust-enshrouded galaxies at redshifts of z ∼ 3 − 4 (24). Electron-based models for the γ ray emission from Mrk 501 (9) predict deviations from a power law in the multi-TeV range even with- out external attenuation and therefore impose an upper limit on the DIRB below the lower limit from faint infrared galaxy counts. If both methods to estimate the DIRB (deviations from a power law spectrum in the multi-TeV range, faint infrared galaxy counts) use correct assumptions, a cutoff in the γ ray spectrum of Mrk 501 must be present in the energy range 10 to 30 TeV. Karl Mannheim 10 References and Notes 1. M.F. Cawley and T.C. Weekes, Exp. Astron. 6, 7 (1995). 2. F. Aharionan, et al. (HEGRA collaboration), Astron. Astrophys. 327, L5 (1997). 3. C.D. Dermer and R. Schlickeiser, Science 257, 1642 (1992). 4. P.O. Lagage and C.J. Cesarsky, Astron. Astrophys. 125, 249 (1983); B. Wiebel-Sooth, P.L. Biermann, and H. Meyer, Astron. Astrophys., in press (1997) (astro-ph/9709253). 5. J.R. Jokipii, Astrophys. J. 313, 842 (1987); J. Bednartz and M. Ostrowski, Mon. Not. R. Astr. Soc. 283, 447 (1996). 6. S.R. Reynolds, Astrophys. J. 459, L13 (1996). 7. O.C. De Jager, et al., Astrophys. J. 457, 253 (1996). 8. K. Koyama, et al., Nature 378, 255 (1995). 9. M. Catanese, et al. (Whipple collaboration), Astrophys. J. 487, L143 (1997); J. Quinn, et al., Astrophys. J. 456, L83 (1996); E. Pian, et al., Astrophys. J. Lett, in press (1997) (astro-ph/9710331). 10. G. Ghisellini, P. Padovani, A. Celotti, and L. Maraschi, Astrophys. J. 407, 65 (1993). 11. F. Krennrich, et al. (Whipple collaboration), Astrophys. J. 481, 758 (1997). 12. D.E. Harris, C.L. Carilli, and R.A. Perley, Nature 367, 713 (1994). 13. C.M. Urry and P. Padovani, Pub. Astro. Soc. Pac. 107, 803 (1995). 14. P.L. Biermann and P.A. Strittmatter, Astrophys. J. 322, 643 (1987). 15. A. Dar and A. Laor, Astrophys. J. 478, L5 (1997). 16. K. Mannheim, P.L. Biermann, and W.M. Krulls, Astron. Astrophys. 251, 723 (1991); K. Mannheim, Astron. Astrophys. 269, 67 (1993). 17. K. Mannheim, S. Westerhoff, H. Meyer, and H.-H. Fink, Astron. Astro- phys. 315, 77 (1996). 18. D.A. Kniffen, et al., Astron. Astrophys. Suppl. 120, 615 (1996); P. Karl Mannheim 11 Padovani, G. Ghisellini, A.C. Fabian, and A. Celotti, Mon. Not. R. Astr. Soc. 260, L21 (1993); C. Impey, Astron. J. 112, 2667 (1996). 19. R.J. Protheroe and P.A. Johnson, Astropart. Phys. 5, 215 (1996). 20. T. Stanev, P.L. Biermann, J. Lloyd-Evans, J.P. Rachen, and A.A. Wat- son, Phys. Rev. Lett. 75, 3065 (1995). 21. J.P. Rachen, T. Stanev, and P.L. Biermann, Astron. Astrophys. 273, 377 (1993); N. Hayashida, et al., Phys. Rev. Lett. 77, 1000 (1996). 22. T.K. Gaisser, F. Halzen, and T. Stanev, Phys. Rep. 258, 173 (1995). 23. D. MacMinn and J.R. Primack, Sp. Sci. Rev. 75, 413 (1996). 24. A. Franceschini, et al., Astron. Astrophys. Suppl. 89, 285 (1991); A. Franceschini, et al., Invited Review in ESA FIRST symposium (ESA SP 401), in press (1997) (astro-ph/9707080); T. Stanev and A. Franceschini, Astrophys. J. Lett., submitted (1997) (astro-ph/9708162). 25. F.W. Stecker and M.A. Malkan, Astrophys. J., in press (1997) (astro- ph/9710072). 26. P. Madau, et al., Mon. Not. R. Astr. Soc. 283, 1388 (1996); B.T. Soifer and G. Neugebauer, Astron. J. 101, 354 (1991). 27. D. Petry, et al., in Proceedings of the 25th International Cosmic Ray Conference, Durban, South Africa, August 1997, eds. P.A. Evanson et al., International Union of Pure and Applied Physics, in press (1998); S.M. Bradbury, et al. (HEGRA collaboration), Astron. Astrophys. 320, L5 (1997). I thank Peter Biermann, Arnon Dar, John Kirk, Hinrich Meyer, Joel 28. Primack, Wolfgang Rhode, Frank Rieger, and the referees for their crit- ical reading and suggestions for improvements of the manuscript. This research was generously supported by the Deutsche Forschungsgemein- schaft under travel grant DFG/Ma 1545/6-1. Karl Mannheim 12 Figure captions Fig.1: Differential TeV spectrum of Mrk 501. Thin solid line: proton blazar model prediction based on archival data from 1995 and earlier (17) taking into account intergalactic attenuation adopting a DIRB spectrum based on a cold+hot dark matter galaxy formation model from (23) and normalized to νIν(25µm) = 1.0 nW m−2 sr−1 (a Hubble constant of H◦ = 75 km s−1 Mpc−1 is assumed throughout the paper). Thin dashed line: same without inter- galactic attenuation. Solid thick lines: model spectrum scaled up by factors 2.5 and 20 corresponding to the increase in the mean sub-TeV flux from 1995 to 1996 and spring of 1997, respectively (9). Open circles: HEGRA (High En- ergy Gamma-Ray Astronomy) CT1 observation March-August 1996 (220h) (27). Solid circles: HEGRA IACT observation March-April 1997 (26.7h) (2) Karl Mannheim 13 x 20 x 2.5 0.0 0.5 log10 E [TeV] 1.0 1.5 Figure 1: 1 0 -1 -2 -3 -4 ] ) s 2 m c V e T ( / 1 1 - 0 1 [ γ F 0 1 g o l
astro-ph/0404536
2
0404
2004-05-17T00:46:24
Relativistic Effects of our Galaxy's Motion on Circles-in-the-sky
[ "astro-ph", "gr-qc" ]
We study the geometric effects of our galaxy's peculiar motion on the circles-in-the-sky. We show that the shape of these circles-in-the-sky remains circular, as detected by a local observer with arbitrary peculiar velocity. Explicit expressions for the radius and center position of such an observed circle-in-the-sky, as well as for the angular displacement of points on the circle, are derived. In general, a circle is detected as a circle of different radius, displaced relative to its original position, and centered at a point which does not correspond to its detected center in the comoving frame. Further, there is an angular displacement of points on the circles. These effects all arise from aberration of cosmic microwave background radiation, exhausting the purely geometric effects due to the peculiar motion of our galaxy, and are independent of both the large scale curvature of space and the expansion of the universe, since aberration is a purely local phenomenon. For a Lorentz-boosted observer with the speed of our entire galaxy, the maximum (detectable) changes in the angular radius of a circle, its maximum center displacement, as well as the maximum angular distortion are shown all to be of order $\beta=(v/c)$ radians. In particular, two back-to-back matching circles in a finite universe will have an upper bound of $2|\beta|$ in the variation of either their radii, the angular position of their centers, or the angular distribution of points.
astro-ph
astro-ph
Relativistic effects of our galaxy's motion on circles -- in -- the -- sky M.O. Calvao∗ Universidade Federal do Rio de Janeiro Instituto de F´ısica, C.P. 68528 21945-972 Rio de Janeiro -- RJ, Brazil G.I. Gomero† Instituto de F´ısica Te´orica Universidade Estadual Paulista Rua Pamplona 145 01405-900 Sao Paulo -- SP, Brazil B. Mota‡ and M.J. Rebou¸cas§ Centro Brasileiro de Pesquisas F´ısicas Departamento de Relatividade e Part´ıculas Rua Dr. Xavier Sigaud 150 22290-180 Rio de Janeiro -- RJ, Brazil (Dated: September 28, 2018) For an observer in the Hubble flow (comoving frame) the last scattering surface (LSS) is well approximated by a two-sphere. If a nontrivial topology of space is detectable, then this sphere intersects some of its topological images, giving rise to circles-in-the-sky, i.e., pairs of matching circles of equal radii, centered at different points on the LSS sphere, with the same pattern of temperature variations. Motivated by the fact that our entire galaxy is not exactly in the Hubble flow, we study the geometric effects of our galaxy's peculiar motion on the circles-in-the-sky. We show that the shape of these circles-in-the-sky remains circular, as detected by a local observer with arbitrary peculiar velocity. Explicit expressions for the radius and center position of such an observed circle-in-the-sky, as well as for the angular displacement of points on the circle, are derived. In general, a circle is detected as a circle of different radius, displaced relative to its original position, and centered at a point which does not correspond to its detected center in the comoving frame. Further, there is an angular displacement of points on the circles. These effects all arise from aberration of cosmic microwave background radiation, exhausting the purely geometric effects due to the peculiar motion of our galaxy, and are independent of both the large scale curvature of space and the expansion of the universe, since aberration is a purely local phenomenon. For a Lorentz-boosted observer with the speed of our entire galaxy, the maximum (detectable) changes in the angular radius of a circle, its maximum center displacement, as well as the maximum angular distortion are shown all to be of order β = (v/c) radians. In particular, two back-to-back matching circles in a finite universe will have an upper bound of 2β in the variation of either their radii, the angular position of their centers, or the angular distribution of points. I. INTRODUCTION Whether the universe is spatially finite and what its size and shape may be are among the fundamental open problems that high precision modern cosmology seeks to resolve. These questions of topological nature have be- come particularly topical, given the wealth of increasingly accurate cosmological observations, especially the recent results from the Wilkinson Microwave Anisotropy Probe (WMAP) experiment [1], which have heightened the in- terest in the possibility of a finite universe. Indeed, re- ported non-Gaussianity in cosmic microwave background (CMB) maps [2], the small power of large-angle fluctua- tions [3], and some features in the power spectrum [2, 3] are large-scale anomalies which have been suggested as potential indication of a finite universe [4] (for reviews see [5]). Given the current high quality and resolution of such maps, the most promising search for cosmic topology is based on pattern repetitions of these CMB anisotropies on the last scattering surface (LSS). If a nontrivial topol- ogy of space is detectable1, then the last scattering sphere intersects some of its topological images giving rise to the so-called circles-in-the-sky. Thus, the CMB temper- ature anisotropy maps will have matched circles: pairs of equal radii circles (centered at different points on the ∗Electronic address: [email protected] †Electronic address: [email protected] ‡Electronic address: [email protected] §Electronic address: [email protected] 1 The extent to which a nontrivial topology may or may not be detected has been discussed in references [6]. LSS sphere) that have the same pattern of temperature variations [7]. These matching circles will exist in CMB anisotropy maps of universes with any detectable non- trivial topology, regardless of its geometry. Therefore, pairs of matched circles may be hidden in CMB maps if the universe is finite, and to observationally probe non- trivial topology on the largest scale available, one needs a statistical approach to scan all-sky CMB maps in order to draw the correlated circles out of them. The circles- in-the-sky method, devised by N. Cornish, D. Spergel and G. Starkman [7] to search for a possible nontrivial topology of the universe, looks for such matching circles through a correlation statistic for sign detection (a func- tion whose peaks indicate matched circles). As originally conceived, the circles -- in -- the -- sky method did not take into account the role of our galaxy's pecu- liar motion. In a recent paper [8], however, this point has been considered, and it has been argued, in a sim- plified context (flat spacetime), that, for any observer moving with respect to the CMB, two effects will take place, namely the circles will be deformed into ovals, and these ovals will be displaced with regard to the corre- sponding circles in the comoving frame. These effects were estimated to be, respectively, of order β 2 and β. We show here that, regardless of any background cur- vature or expansion, the shape of these circles -- in -- the -- sky, as locally detected by an observer in motion relative to the comoving one, remains circular. We derive ex- plicit expressions for the radius and center position of such an observed circle-in-the-sky, as well as a formula for the angular displacement of points on the circle. In general, a circle is detected as a circle of different radius, displaced relative to its original position, and centered at a point which does not correspond to its detected cen- ter in the comoving frame. Further, there is an angular displacement of points on the circles. These effects arise all from aberration of CMBR, and exhaust the purely geometric effects due to our galaxy's peculiar motion on circles-in-the-sky. We also estimate the maximum val- ues of these effects considering the peculiar motion of our galaxy. In particular, for two back-to-back matching circles in a finite universe the upper bounds in the vari- ation of either their radii, the angular position of their center or in the angular distribution of points are all of order 2β ≃ 2.46 × 10−3 radians. Although these effects are still below WMAP's angular resolution, we show that they are relevant for future CMB missions like the Planck satellite. II. MAIN RESULTS We begin by recalling an old beautiful result by R. Pen- rose [9], according to which, if an observer O detects, in his sky-sphere, the shape of an object as circular, then for any other observer O′ locally coinciding with and in mo- tion relative to O, the detected shape will remain circular. The gist of the reasoning is based on the relativistic aber- 2 ration and the stereographic projection properties, as de- scribed in more detail by W. Rindler [10]: the outline of a spherical object, projected onto the local celestial sphere ("sky") of a given observer O is certainly circular and, by means of a suitable stereographic projection, is mapped onto a circle in the projection plane ("screen"). Now, the aberration formula just transforms this screen for O onto a corresponding screen for O′ which is merely globally expanded (or contracted) by a factor [(1 − β)/(1 + β)]1/2; by the corresponding inverse stereographic projection, we thus arrive at the above mentioned Penrose's result. As a consequence, circles-in-the-sky as detected by the ob- server O will also be detected as circles-in-the-sky by the observer O′. The circles will, in general, differ both in position and in size as we shall discuss in detail in what follows. Let P µ = (hν/c)(1, −bn) be the 4-momentum of an in- coming photon in a direction bn and with frequency ν as detected by an observer O in the Hubble flow (comov- ing observer). For another observer O′ who moves with velocity ~β relative to this comoving observer, and whose spatial position coincides with that of O at the time they measure the CMB, the photon will be detected as in- coming in a different direction bn′, and with a distinct frequency ν′. Clearly, the 4-momentum of this photon for the observer O′ is given by the Lorentz transforma- tion (1) (2) P ′µ = Λµ σP σ , where σ =(cid:18) γ −γ ~β Λµ −γ ~βT I + (γ − 1)bβ bβT (cid:19) . Here I stands for the 3-dimensional identity matrix, γ = (1 − β 2)−1/2, ~βT it the transpose of ~β, and bβ = ~β/β. Thus we readily obtain ν′ = γ(cid:16)1 + ~β ·bn(cid:17) ν , bn′ = bn +h(γ − 1)bβ ·bn + γβibβ γ(cid:16)1 + ~β ·bn(cid:17) (3) (4) . Equation (3) gives the Doppler effect, i.e. a shift in frequency due to our galaxy's peculiar motion, respon- sible for the large dipole moment in the CMB tempera- ture anisotropies and for tiny spectral distortions in the corresponding maps. Equation (4), on the other hand, expresses the relativistic aberration of light, which gives rise to the position displacement and re-scaling as well as the angular displacement of points of the circles-in-the- sky due to the motion of O′ relative to O. We first use (4) to determine the center and radius of a Lorentz-boosted circle-in-the-sky. Without loss of generality, let us choose the common axes z and z′ such that their positive direction coincides with the direction of the velocity ~β. Thus, from (4), the transformation of the direction of the incoming photon, (θ, ϕ) 7→ (θ′, ϕ′), reduces to cos θ′ = β + cos θ 1 + β cos θ and ϕ′ = ϕ , (5) where θ and ϕ are the usual spherical coordinates. Let bqc = (θc, ϕc) and ρ be, respectively, the center and radius of a circle C(bqc, ρ) in the comoving frame. We note that the direct use of (5) to the center bqc of C does not furnish, in general, the center bq′ circle C ′(bq′ to calculate the expressions for the center bq′ c of the Lorentz-boosted c, ρ′), because stereographic projection of a cir- cle C is a circle whose center is not, in general, the stereo- graphic projection of the center of C. Therefore, in order c, ϕ′ c) and the radius ρ′ of the Lorentz-boosted circle C ′ we use (5) to transform the top and bottom points of the circle C, given by (θc − ρ, ϕc) and (θc + ρ, ϕc), respec- tively. As ϕ′ = ϕ, obviously one has ϕ′ c = ϕc, and the transformed points remain, respectively, the top and bot- tom points of the Lorentz-boosted circle C ′. From (5) we have c = (θ′ cos(θ′ c − ρ′) = β + cos(θc − ρ) 1 + β cos(θc − ρ) (6) for the top point, and a similar expression for (θ′ c + ρ′). Now using some elementary trigonometric relations, after some calculation we obtain sin θ′ c = 1 γ M sin θc , and sin ρ′ = 1 γ M sin ρ , (7) where M =p[1 + β cos(θc + ρ)][1 + β cos(θc − ρ)] . (8) We emphasize that, by using (5) and (7), it can readily be verified that the circle-in-the-sky C(bqc, ρ) transforms into the circle-in-the-sky C ′(bq′ c, ρ′), i.e., the circle equa- tion bq ·bqc = cos ρ (9) is invariant under the transformation given by (5) and (7). circle (cf. eq.(12) below). In (9) bq stands for an arbitrary point in the Now, for non-relativistic velocities (β ≪ 1), at first order in β we have θ′ c = θc − β sin θc cos ρ , ρ′ = ρ − β cos θc sin ρ . (10) (11) It follows from (10) that for a given circle of radius ρ the displacement ∆θc = θ′ c − θc of its center is max- imum for θc = π/2, i.e. when the vector position of its center in the comoving frame is orthogonal to the boost- velocity ~β. In this case the angular radius ρ remains unchanged as detected by O′. The upper bound of ∆θc clearly is β. Thus, two back-to-back circles-in-the-sky 3 in a universe with a nontrivial topology will have a maxi- mum displacement of 2β in the position of their centers. Figure 1 shows, for a fixed β, the behavior of ∆θc as a function of θc for circles of different radii. It is apparent from this figure, that the maximum absolute value of the angular displacement of the center of the circles depends on their radius but occurs for θc = π/2, while for a circle of angular radius ρ = π/2 there is no change in the posi- tion of the center no matter what the center position is. ρ= 0o ρ= 45o ρ= 60o ∆θc 0.07 0.06 0.05 0.04 0.03 0.02 0.01 30 60 ρ= 90o 90 120 150 180 θc FIG. 1: The behavior of displacement ∆θc of the position of the center for different radii of the circles. In this figure β = 1.23 × 10−3. On the other hand, for a given circle C of angular ra- dius ρ the change ∆ρ = ρ′ − ρ is maximum when the direction of vector position of the center of C in the co- moving frame is the same direction of ~β, i.e. θc = 0 or θc = π. In each of these cases the position of the center of the circle as detected by O′ remains fixed. We note that when θc = 0 the radius of the circle decreases as detected by the observer O′, while for θc = π it increases (see also figure 3). The upper bound for ∆ρ clearly is again β. Figure 2 shows the behavior of ∆ρ as a func- tion of ρ for distinct positions of the centers θc. It is clear from this figure that circles whose vector position of the center is orthogonal to the velocity ~β do not change the radii, while for circles whose vector position is parallel to (antiparallel) β we have a maximum absolute variation of ρ for any given fixed circle. Clearly for 0◦ ≤ θc < 90◦, ∆ρ is a decreasing function of ρ, while for 90◦ < θc ≤ 180◦ it is an increasing function of ρ. From (11) we have that for θc = 90◦ the radius re- mains unchanged. The upper bound for ∆ρ obviously is β. Thus, two back-to-back correlated circles in a fi- nite universe with a detectable topology will have a upper bound of 2β in the variation of the radii as detected by the moving observer O′. Figure 3 shows the changes in the radii and in the positions of the center of the circles in three instances. The continuous and dashed circles indicate circles-in-the- sky as detected, respectively, by the observers O and O′. The circles C2 and C ′ 2 correspond to a case of maximum change in the position of the center. Clearly in this case the radius ρ remains unchanged at first order in β. The circles C1 and C ′ 1 correspond to a case in which the radius of the circle decreases, while C3 and C ′ 3 shows a case when ∆ρ 0.06 0.04 0.02 -0.02 -0.04 -0.06 15 30 45 60 75 θc = 180o θc =135o θc =90o 90 ρ θc = 45o θc = 0o 4 plement of the projection of bq onto bqc. We take ~p0, the orthogonal complement of the top point bq0 = (θc − ρ, ϕc) of C, as a reference from which we measure the angle α defined by cos α = ~p · ~p0 sin2 ρ , (13) since clearly ~p = ~p0 = sin ρ. From (12) one easily has FIG. 2: The behavior of displacement ∆ρ of the radius of the circles as a function of ρ for different locations of the center of the circles. In this figure β = 1.23 × 10−3. cos α = cos θ sin θc − sin θ cos θc cos(ϕ − ϕc) sin ρ , (14) the radius increases. In these two last cases the positions of the center of the circles remain unchanged. The pairs of matching circles (C1, C3) and (C ′ 3) illustrate the instance of maximum change in back-to-back correlated circles-in-the-sky. 1, C ′ C'1 C1 ρ'1 ρ 1 ρ'1< ρ 1 ρ ' 2 β C ' 2 ρ'2= ρ 2 C2 ρ 2 ρ'3> ρ 3 C'3 C3 ρ'3 ρ 3 FIG. 3: A schematic illustration of the changes in the radii and in the positions of the center of the circles. The continu- ous and dashed circles indicate circles-in-the-sky as detected, respectively, by the observers O and O′. The circles C2 and C ′ 2 illustrate the maximum change in the position of the cen- ter. In this case the radius remains the same. The circles C1 and C ′ 1 represent the instance in which the radius of the cir- cle has a maximum decrease, while C3 and C ′ 3 illustrate the case when the radius has a maximum increase. The set of cir- cles {(C1, C3), (C ′ 3)} stands for the situation of maximum change in back-to-back matching circles-in-the-sky. 1, C ′ To fully characterize all geometric effects which arise from aberration of CMB on circles-in-the-sky we shall consider now the angular displacement of points on a circle as detected by the observers O and O′. To this end, we decompose a point bq = (θ, ϕ) on a circle C(bqc, ρ) as where ~p is on the plane of the circle, which is orthogonal to bqc = (θc, ϕc). The vector ~p is the orthogonal com- bq = ~p +bqc cos ρ , (12) or using the circle equation (9) to eliminate the depen- dence on ϕ − ϕc, cos α = cos θ − cos θc cos ρ sin θc sin ρ . (15) To calculate cos α′ it is more convenient to substitute (5) and (7) into the analog of (14), for cos α′ (which clearly must hold) and then use (9) and (15) to simplify the expression to cos α′ = cos α +(cid:18) 1 + β cos θ(cid:19) sin ρ sin θc sin2 α , β (16) where cos θ = cos ρ cos θc + sin ρ sin θc cos α. At first order in β equation (16) becomes α′ = α − β sin ρ sin θc sin α . (17) It follows from (17) that for a given circle of radius ρ the angular displacement ∆α is maximum for θc = π/2 and α = ±π/2. The upper bound for ∆α clearly is β and occurs when simultaneously θc = π/2, ρ = π/2 and α = π/2. Here again, two back-to-back matching circles in a multiply connected universe will have a upper bound of 2β in the variation of the angular displacement. Fig- ure 4 shows the behavior of ∆α as a function of α for different values of the radii and location of the center of circles. It is clear from this figure that for any given circles center at θc 6= 0◦ the value of maximum angular distortion depend on the radii and centers of the circles but takes place for α = ±π/2, while for circles with ρ whose vector position of the centers are parallel (or an- tiparallel) to ~β (θc = 0◦ or θc = π) there is no angular distortion. Figures 5 and 6 illustrate the angular distortion effect. In these figures it is shown how a set of points uniformly comoving observer O will be detected by a locally coin- distributed along the circle C(bqc, ρ) as detected by the ciding observer O′ in the transformed circle C ′(bq′ c, ρ′). III. CONCLUDING REMARKS Motivated by the fact that our entire galaxy is not in the Hubble flow, we have studied all geometric effects θc = 90o , ρ = 90o θc = 90o , ρ = 45o θc = 45o , ρ = 45o ∆α 0.07 0.06 0.05 0.04 0.03 0.02 0.01 θc = 0o or ρ = 0o 45 225 90 270 135 315 180 α 360 FIG. 4: The behavior of angular displacement ∆α for circles of different radii and centers. In this figure β = 1.23 × 10−3. z β θ'c θ c ϕ = ϕ ' x ρ' ρ y FIG. 5: A schematic illustration of the angular displacement effect. It shows how a set of points homogeneously distributed on the solid line circle C(bqc, ρ) as detected by an observer O, is detected by the observer O′ in the new dashed circle C ′(bq′ c, ρ′). 1 0.5 α 1 0.5 α' -1 -0.5 0.5 1 -1 -0.5 0.5 1 -0.5 -1 -0.5 -1 FIG. 6: Points regularly distributed along a circle (left), as detected by observer O, are distorted by aberration when de- tected by observer O′ (right). The points lying on the merid- ian (α=0 and α = 180◦) are unaffected by angular distortion. In this figure the exact expression (16) was used with β = 0.8. due to our galaxy's peculiar motion on the circles-in-the- sky. As a consequence of Peronse's result [9] we have 5 shown that circles-in-the-sky as detected by a comoving observer O will also be detected as circles-in-the sky by any observer O′ locally coinciding with O at the time they measure the CMBR. In general, a circle for O will be detected by O′ as a circle of different radius, displaced relative to its original position, and centered at a point which does not correspond to its detected center in the comoving frame. We have derived closed explicit expressions for the ra- dius, ρ′, the center position, (θ′ c), and the angular displacement α′ of points on the circles as detected by the Lorentz-boosted observer O′. We have also plotted figures to illustrate several instances of these effects. c, φ′ The maximum displacement in either radius or center or angular position of points on a circle is ±β. From the value β = 1.23 × 10−3 obtained from the dipole amplitude in the CMB spectrum, we have that the max- imum displacement for each of these effects individually is ≃ ±0.07◦ = 4.2′. Thus, two antipodal circles whose center vector positions are parallel (antiparallel) to ~β will have a difference in radii of 0.14◦, as detected by O and O′. This is below the current WMAP resolution (at best 0.25◦ = 15′) [1], but close to Planck's (at least 0.16◦ = 10′) [11] and other more accurate forthcoming missions' resolution. In Ref. [8], in a simplified flat spacetime, the spatial positions, relative to a "moving" frame, of the events de- fined by the intersection of the copies of the LSS spheres were determined; of course, these events are simultaneous in the CMB frame and, ipso facto, are non-simultaneous in the "moving" one. The specific procedure of collecting the spatial positions of these non-simultaneous events de- fines, in the "moving" frame, an oval figure, whose diam- eter along the boost direction is Lorentz-dilated, whereas its perpendicular diameter remains unchanged. Briefly, it was shown that a (spatial) circle in the CMB frame, as a geometric figure, is formally transformed into a (spa- tial) oval in the "moving" frame of the underlying flat spacetime. The approach we adopted to the effects of our galaxy's peculiar motion on the circles-in-the-sky is realistic and complete, since it does not rely on such a flat spacetime, and takes into account that a typical ob- servation of the CMBR records simultaneously incoming light rays in an essentially infinitesimal detector used by a local observer, thus implying their projection onto the sky-sphere. We note that the aberration of CMBR might be of a more general interest in the analysis of maps of CMB temperature anisotropies. Indeed, consider the celestial sphere pixelized by a comoving observer O and consider another observer O′ coinciding with O and with relative velocity ~β along the positive z direction. The coordinates (θi, ϕi) of the i -- th pixel in the comoving frame are trans- formed to the moving frame according to (5), which, at first order in β, simply reads θ′ i = θi − β sin θi ; ϕ′ i = ϕi . (18) This displacement of the centers of the pixels gives rise to a distortion on the temperature variation pattern in CMB maps, which might be relevant for future missions. Finally, we emphasize that all these effects arise from aberration of the CMBR, and exhaust the purely geo- metric effects due to the motion of our galaxy (or CMBR detectors). It is again worth noting that, since aberration of light is a purely local phenomenon, the effects con- sidered in this work depend neither on the background curvature of space nor on the expansion of the universe. Acknowledgments We thank CNPq and FAPESP (contract 02/12328-6) for the grants under which this work was carried out. 6 [1] C.L. Bennett et al. , Astrophys. J. 583, 1 (2003); G. Hinshaw et al. , Astrophys. J. Suppl. 148, 135 (2003); C.L. Bennett et al. , Astrophys. J. Suppl. 148, 1 (2003). [2] A. de Oliveira-Costa, M. Tegmark, M. Zaldarriaga, A. Hamilton, Phys. Rev. D 69 063516 (2004); H.K. Eriksen, F.K. Hansen, A.J. Banday, K.M. Gorski, P.B. Lilje, Asymmetries in the CMB anisotropy field , astro-ph/0307507. [3] D.N. Spergel et al. , Astrophys. J. Suppl. 148,175 (2003). [4] J.-P. Luminet, J. Weeks, A. Riazuelo, R. Lehoucq and J.-P. Uzan, Nature 425, 593 (2003). [5] G.F.R. Ellis, Gen. Rel. Grav. 2 7 (1971); M. Lachi`eze-Rey and J.-P. Luminet, Phys. Rep. 254, 135 (1995); G.D. Starkman, Class. Quantum Grav. 15, 2529 (1998); J. Levin, Phys. Rep. 365, 251 (2002); M.J. Rebou¸cas and G.I. Gomero, astro-ph/0402324. To appear in Braz. J. Phys. (2004). [6] G.I. Gomero, M.J. Rebou¸cas and R. Tavakol, Class. Quantum Grav. 18, 4461 (2001); G.I. Gomero, M.J. Re- bou¸cas and R. Tavakol, Class. Quantum Grav. 18, L145 (2001); G.I. Gomero, M.J. Rebou¸cas and R. Tavakol, Int. J. Mod. Phys. A 17, 4261 (2002); J.R. Weeks, R. Lehoucq, J.-P. Uzan, Class. Quant. Grav. 20, 1529 (2003); J.R. Weeks, Mod. Phys. Lett. A 18, 2099 (2003); G.I. Gomero and M.J. Rebou¸cas, Phys. Lett. A 311, 319 (2003); B. Mota, M.J. Rebou¸cas and R. Tavakol, Class. Quantum Grav. 20, 4837 (2003). [7] N.J. Cornish, D. Spergel and G. Starkman, Class. Quan- tum Grav. 15, 2657 (1998). [8] J. Levin, Missing Lorenz -- boosted Circles-in-the-sky, astro-ph/0403036. [9] R. Penrose, Proc. Cambridge Philos. Soc. 55, 137 (1959). [10] W. Rindler, Essential Relativity, revised second edition, Chapter 3, Springer-Verlag, Heidelberg (1979). [11] V. Stolyarov et al.,Mon. Not. Roy. Astron. Soc. 336, 97 (2002); http://www.rssd.esa.int
astro-ph/0702075
1
0702
2007-02-02T17:19:44
Newborn Magnetars as sources of Gravitational Radiation: constraints from High Energy observations of Magnetar Candidates
[ "astro-ph" ]
Soft Gamma Repeaters and the Anomalous X-ray Pulsars are believed to contain slowly spinning "magnetars". The enormous energy liberated in the 2004 Dece 27 giant flare from SGR 1806-20, together with the likely recurrence time of such events, points to an internal magnetic field strength ~ 10^{16} G. Such strong fields are expected to be generated by a coherent alpha-Omega dynamo in the early seconds after the Neutron Star formation, if its spin period is of a few milliseconds at most. A substantial deformation of the NS is caused by such fields and a newborn millisecond-spinning magnetar would thus radiate for a few days a strong gravitational wave signal. Such a signal may be detected with Advanced LIGO-class detectors up to the distance of the Virgo cluster, where ~ 1 magnetar per year are expected to form. Recent X-ray observations reveal that SNRs around magnetar candidates do not show evidence for a larger energy content than standard SNRs (Vink & Kuiper 2006). This is at variance with what would be expected if the spin energy of the young, millisecond NS were radiated away as electromagnetic radiation andd/or relativistic particle winds and, thus, transferred quickly to the expanding gas shell. We show here that these recent findings can be reconciled with the idea of magnetars being formed with fast spins, if most of their initial spin energy is radiated thorugh GWs. In particular, we find that this occurs for essentially the same parameter range that would make such objects detectable by Advanced LIGO-class detectors up to the Virgo Cluster.
astro-ph
astro-ph
Astrophysics and Space Science manuscript No. (will be inserted by the editor) 7 0 0 2 b e F 2 1 v 5 7 0 2 0 7 0 / h p - o r t s a : v i X r a S. Dall'Osso · L. Stella Newborn Magnetars as sources of Gravitational Radiation: constraints from High Energy observations of Magnetar Candidates Received: date / Accepted: date Abstract Two classes of high-energy sources, the Soft Gamma Repeaters and the Anomalous X-ray Pulsars are believed to contain slowly spinning "magnetars", i.e. neutron stars the emission of which derives from the re- lease of energy from their extremely strong magnetic fields (> 1015 G). The enormous energy liberated in the 2004 December 27 giant flare from SGR 1806-20 (∼ 5 × 1046 erg), together with the likely recurrence time of such events, points to an internal magnetic field strength of ≥ 1016 G. Such strong fields are expected to be generated by a coherent α − Ω dynamo in the early seconds after the Neutron Star (NS) formation, if its spin period is of a few milliseconds at most. A sub- stantial deformation of the NS is caused by such fields and, provided the deformation axis is offset from the spin axis, a newborn millisecond-spinning magnetar would thus radiate for a few days a strong gravitational wave signal the frequency of which (∼ 0.5 − 2 kHz range) decreases in time. This signal could be detected with Advanced LIGO-class detectors up to the distance of the Virgo cluster, where ≥ 1 yr−1 magnetars are ex- pected to form. Recent X-ray observations revealed that SNRs around magnetar candidates do not appear to have received a larger energy input than in standard SNRs (Vink & Kuiper 2006). This is at variance with what would be expected if the spin energy of the young, mil- lisecond NS were radiated away as electromagnetic ra- diation andd/or relativistic particle winds. In fact, such energy would be transferred quickly and efficiently to the expanding gas shell. This may thus suggest that mag- netars did not form with the expected very fast initial spin. We show here that these findings can be reconciled S. Dall'Osso INAF-Osservatorio Astronomico di Roma via di Frascati 33, 00040, Monteporzio Catone (Roma) Tel.: +39 06 94 28 64 37 E-mail: [email protected] L. Stella INAF-Osservatorio Astronomico di Roma via di Frascati 33, 00040 Monteporzio Catone (Roma) Tel.: +39 06 94 28 64 36 E-mail: [email protected] with the idea of magnetars being formed with fast spins, if most of their initial spin energy is radiated thorugh GWs. In particular, we find that this occurs for essen- tially the same parameter range that would make such objects detectable by Advanced LIGO-class detectors up to the Virgo Cluster. If our argument holds for at least a fraction of newly formed magnetars, then these objects constitute a promising new class of gravitational wave emitters. Keywords gravitational waves - stars: magnetic fields - stars: neutron - stars: individual: SGR 1806-20) PACS 97.60.Jd · 97.60.Bw · 04.30.Db · 95.85.Sz 1 Introduction The Soft Gamma Repeaters, SGRs, and the Anomalous X-ray Pulsars, AXPs, have a number of properties in common (Mereghetti & Stella 1995; Kouveliotou et al. 1998; Woods & Thompson 2004). They have spin peri- ods of ∼ 5 ÷10 s, spin-down secularly with ∼ 104 ÷105 yr timescale, are isolated and in some cases associated to su- pernova remnants with ∼ 103 ÷ 104 yr ages. Rotational energy losses are 10 ÷ 100 times too low to explain the ∼ 1034 ÷ 1035 erg/s persistent emission of these sources. Both AXPs and SGRs have periods of intense activity during which recurrent, subsecond-long bursts are emit- ted (peak luminosities of ∼ 1038 ÷ 1041 erg/s). The ini- tial spikes of giant flares have comparable duration but 3 to 6 orders of magnitude larger luminosity. Giant flares are rare, only three have been observed in about 30 yr of monitoring. Given the highly super-Eddington lumi- nosities of recurrent bursts and, especially, giant flares, accretion models are not viable. In the magnetar model, SGRs and AXPs derive their emission from the release of the energy stored in their ex- tremely high magnetic fields (Duncan & Thompson 1992; Thompson & Duncan 1993, 1995, 1996, 2001). This is the leading model for interpreting the unique features of these sources. According to it, a wound-up, mainly 2 Dall'Osso et al. A 2004 Dec 27-like event in the Galaxy could not be missed, whereas several systematic effects can reduce the chances of detection from large distances (see e.g. the discussion in Lazzati et al. (2005); Nakar et al. (2005)). Rather than regarding the 2004 Dec 27 event as statis- tically unlikely, one can thus evaluate the chances of a recurrence time of hundreds of years, given the occur- rence of the 2004 Dec 27 hyperflare. We estimate that, having observed a powerful giant flare in our galaxy in ∼ 30 yr of observations, the Bayesian probability that the galactic recurrence time is τ > 600 yr is ∼ 10−3, whereas the 90% confidence upper limit is τ ∼ 60 yr. We thus favor smaller values and assume in the following τ ∼ 30 yr. In ∼ 104 yr (that we adopt for the SGR lifetime), about 70 very powerful giant flares should be emitted by an SGR, releasing a total energy of ∼ 4 × 1048 erg. We note that if the giant flares' emission were beamed in a fraction b of the sky (and thus the energy released in in- dividual flares a factor of b lower), the recurrence time would be a factor of b shorter. Therefore the total release of energy would remain the same. If this energy origi- nates from the magnetar's internal magnetic field, this must be ≥ 1015.7 G (Stella et al. 2005; Terasawa et al. 2005). This value should be regarded as a lower limit. Firstly, the magnetar model predicts a conspicuous neu- trino luminosity from ambipolar diffusion-driven field de- cay, an energy component that is not available to flares. Including this, we estimate that the limit above increases by ∼ 60% and becomes B ≥ 1015.9 G. Secondly, ambipo- lar diffusion and magnetic dissipation should take place at a faster rate for higher values of the field (Thompson & Duncan 1996). Therefore estimates of the internal B-field based on present day properties of SGRs likely underestimate the value of their initial magnetic field. Very strong toroidal B fields are expected to be gener- ated inside a differentially rotating fast spinning neutron star, subject to vigorous neutrino-driven convection in- stants after its formation (Duncan & Thompson 1992). A field of several ×1016 G can be generated in magne- tars that are born with spin periods of a few milliseconds (Thompson & Duncan 1993). As discussed by Duncan (1998), values up to ∼ 1017 G cannot be ruled out. In the following we explore the consequences of these fields for the generation of gravitational waves from new- born magnetars. We parametrize their (internal) toroidal field with Bt,16.3 = Bt/2 × 1016 G, (external) dipole field with Bd,14 = Bd/1014 G and initial spin period with Pi,2 = Pi/(2 ms). toroidal magnetic field characterizes the neutron star in- terior (B > 1015 G). The emerged (mainly poloidal) field makes up the neutron star magnetosphere; dipole strenghts (Bd ∼ few ×1014 G) are required to generate the observed spin- down(Thompson & Duncan 1993; Thompson & Murray 2001). Impulsive energy is fed to the neutron star magne- tosphere through Alfv´en waves driven by local "crustquakes" and producing recurrent bursts with a large range of am- plitudes. Giant flares likely originate in large-scale rear- rangements of the toroidal inner field or catastrophic in- stabilities in the magnetosphere (Thompson & Duncan 2001; Lyutikov 2003). Most of this energy breaks out of the magnetosphere in a fireball of plasma expanding at relativistic speeds which produces the initial spike of giant flares. The oscillating tail that follows this spike, displaying many tens of cycles at the neutron star spin, is interpreted as due to a "trapped fireball", which remains anchored inside the magnetosphere (the total energy re- leased in this tail is ∼ 1044 erg in all three events detected so far, comparable to the energy of a ∼ 1014 G trapping magnetospheric field). 2 The 2004 December 27 Event and the Internal Magnetic field of Magnetars The 2004 December 27 giant flare from SGR1806-20 pro- vides a new estimate of the internal field of magnetars. About 5 × 1046 erg were released during the ∼ 0.6 s long initial spike of this event (Terasawa et al. 2005; Hurley et al. 2005). This is more than two decades higher than the en- ergy of the other giant flares observed so far, the 1979 March 5 event from SGR 0526-66 (Mazets et al. 1979) and the 1998 August 27 event from SGR 1900+14 (Hurley et al. 1999; Feroci et al. 1999). Only one such powerful flare has been recorded in about 30 yr of monitoring of the ∼ 5 known magnetars in SGRs. The recurrence time in a single magnetar implied by this event is thus about ∼ 150 yr. The realisation that powerful giant flares could be observed from distances of tens of Mpc (and thus might represent a sizeable fraction of the short Gamma Ray Burst population) motivated searches for 2004 Dec 27- like events in the BATSE GRB database (Lazzati et al. 2005; Popov & Stern 2005). The upper limits on the re- currence time of powerful giant flares obtained in these studies range from τ ∼ 130 to 600 yr per galaxy, i.e. ∼ 4 to 20 times longer than inferred above. Therefore these authors conclude that on 2004 Dec 27 we have witnessed a "statistically unlikely" event. On the other hand Tanvir et al. (2005) find that the location of 10- 25% of the short GRBs in the BATSE catalogue corre- lates with the position of galaxies in the local universe (< 110 Mpc), suggesting that a fraction of the short GRBs may originate from a population of powerful Gi- ant Flare-like events. 3 Magnetically-Induced Distortion and Gravitational Wave Emission The possibility that fast-rotating, magnetically-distorted neutron stars are conspicuous sources of gravitational ra- diation has been discussed by several authors (Bonazzola & Marck GWs from newborn magnetars 3 1994; Bonazzola & Gourgoulhon 1996). More recently, work has been carried out in the context of the mag- netar model, for internal magnetic fields strengths of ∼ 1014 − 1016 G (Konno, Obata, Kojima 2000; Palomba 2001; Cutler 2002). In the following we show that for the range of magnetic fields discussed in Section 2, newly born, millisecond spinning magnetars are conspicuous sources of gravitational radiation that will be detectable up to Virgo cluster distances (Stella et al. 2005). The anisotropic pressure from the toroidal B-field de- forms a magnetar into a prolate shape, with ellipticity ǫB ∼ −6.4 × 10−4(< B 2 t,16.3 >), where the brackets indicate a volume-average over the entire core (Cutler 2002). As long as the axis of the magnetic distortion is not aligned with the spin axis, the star's rotation will cause a periodic variation of the mass quadrupole mo- ment, in turn resulting in the emission of gravitational waves, GWs, at twice the spin frequency of the star. Free precession of the ellipsoidal NS is also excited and, as shown by Mestel & Takhar (1972) and Cutler (2002), its viscous damping drives the symmetry axis of the mag- netic distortion orthogonal to the spins axis, if the el- lipsoid is prolate, i.e. if the magnetic field is toroidal. Therefore, viscous damping of free precession in newly born magnetars leads to a geometry that maximizes the time-varying mass quadrupole moment, and GW emis- sion accordingly. However, the power emitted in GWs scales as ∝ P −6. Therefore the GW signal, for a given toroidal B-field, de- pends critically on the initial value and early evolution of the spin period. The spin evolution of a newborn mag- netar is determined by angular momentum losses from GWs, electromagnetic dipole radiation and relativistic winds. According to Thompson et al (2004), the latter mechanism is negligible except for external dipole fields < (6 ÷ 7) × 1014 G and we will neglect it here. The spin evolution of a newborn magnetar under the combined effects of GW and electromagnetic dipole ra- diation is given by ω = −Kdω3 − Kgwω5 , (1) dR6)/(6 Ic3) and Kgw = (32/5)(G/c5)Iǫ2 where ω = 2π/P is the angular velocity, Kd = (B 2 B, with R the neutron star radius, G the gravitational constant and c the speed of light. This gives a spin-down timescale of τsd ≡ ω 2 ω i,2(cid:0)B 2 ≃ 10 P 2 d,14 + 1.15B 4 t,16.3P −2 d (2) i,2 (cid:1)−1 The condition for the newly formed magnetar to become an orthogonal rotator before loosing a significant fraction of its initial spin energy is (Stella et al. 2005): τsd τort ≃ 26 B 2 t,16.3 B 2 d,14P −1 i,2 + 1.15 B 4 t,16.3P −3 i,2 > 1 . (3) If condition (3) is met, the magnetar quickly becomes a maximally efficient GW emitter, while its spin period is 2 B 2 t,16.3 , 20 P −2 still close to the initial one. In this case, the instanta- neous signal strain can be expressed as: h ∼ 3 × 10−26d−1 where the distance d20 = d/(20 Mpc) is in units of the Virgo Cluster distance and the angle-averaged strain is that given by Ushomirsky et al (2000). We estimate the characteristic amplitude, hc = hN 1/2, where N ≃ τsd/Pi is the number of cycles over which the signal is observed. Using equation (4) we obtain: (4) hc ≃ 6 × 10−22 B 2 t,16.3 (5) 1 2 d20P 3/2 i,2 (cid:16)B 2 d,14 + 1.15B 4 i,2 (cid:17) t,16.3P −2 Under the conditions discussed above, strong GW losses are not quenched immediately after the magnetar birth but rather extend in time, typically from days to a few weeks, before fading away as a result of the star spin- down. The characteristic amplitude in eq. (5) is within reach of GW interferometers of the Advanced LIGO class. In order to assess the detectability of these GW signal we compute the optimal (matched-filter) signal-to-noise ratio for a signal sweeping the 500 Hz - 2 kHz band by us- ing the current baseline performance of Advanced LIGO (details are given in Stella et al. (2005)). Fig. 1 shows lines of constant S/N for a source at d20 = 1 and se- lected values of the initial rotation period (Pi = 1.2, 2 and 2.5 ms) in the (Bt, Bd) plane: GWs from newborn magnetars can produce S/N > 8 for Bt ≥ 1016.5 G and Bd ≤ 1014.5 G. This is the region of parameter space that offers the best prospects for detection as we now discuss. Matched-filtering represents the optimal de- tection strategy for long-lived, periodic signals, also in those cases where their frequency is slowly evolving over time. This process involves the correlation of the data stream with a discrete set of template signals that probe the relevant space of unknown parameters: the sky po- sition (2 extrinsic parameters) and the 2 intrinsic pa- rameters that control the evolution of the GW phase (see Eq.(1)). Template spacing in the parameter space is chosen appropriately, so as to reduce to a value less than, say, 10% (depending on the sensitivity one wants to achieve) the fraction of the intrinsic signal-to-noise ra- tio that is lost in the cross-correlation. Dall'Osso & Re (2006) have recently investigated in detail a matched- filtering (coherent) search strategy for the expected sig- nal from newly formed magnetars in the Virgo cluster. Given the two (uncorrelated) parameters involved in the signal frequency evolution, this approach implies an un- affordable computational cost. A huge number of tem- plates ∼ 1018 would be required to cover the relevant parameter space (the plane Bd vs. Bt) with a sufficiently fine grid that the intrinsic signal-to-noise ratio would not be too degraded. Although this calculation is rather ide- alized, the resulting number of templates is so large that no realistic calculation could decrease it by the several orders of magnitude needed to make the search feasi- ble. The search for this new class of signals represents 4 Dall'Osso et al. Fig. 1 Lines of constant S/N for selected values of the initial spin period in the internal toroidal magnetic field, Bt, and external dipole field, Bt, plane for a source at the distance of the Virgo Cluster (d20 = 1). Solid, dashed and dotted curves correspond to an initial spin period of Pi =1.2, 2 and 2.5 ms, respectively. The calculations take into account the time required for the toroidal magnetic field axis to become orthogonal to the spin axis. Note that according to Thompson et al (2004), strong angular momentum losses by relativistic winds set in and dominate the spin down for Bd > 6 ÷ 7 × 1014 G; the curves for such values of Bd should thus be treated with caution. (2006) deduced from their measurements an initial spin Pi ≥ (5 ÷ 6) ms for the above mentioned sources, assum- ing that all the spin-down energy is emitted through elec- tromagnetic radiation and/or particle winds, and thus absorbed by the surrounding ejecta in the early days of spindown. Their limit period is long enough to rise a se- rious question as to the viability of the α − Ω dynamo scenario for generating the large-scale magnetic fields of magnetars. Models that do not rely upon very short spin periods at birth, such as the flux-freezing scenario sug- gested by Ferrario & Wickramasinghe (2006), would be favored by these results. However, given the possibility that newly formed mag- netars be strong GW emitters in their early days, we show that the results by Vink & Kuiper (2006) can be accounted for within this framework. Most of a magne- tar's initial spin energy could indeed be released through GWs, without being absorbed by the expanding remnant shell. Therefore, the results of X-ray studies can be used within our model to constrain the initial combination of Pi, Bd, Bt. however a challenge that must be investigated in greater depth. We are currently studying a hierarchical approach to the problem, a process where coherent and incoherent stages of the search are alternated as to reduce the com- putational requirements by a large factor, at the price of a relativeley modest loss in sensitivity to the signal. 4 Observational Constraints on GW emission from newly formed magnetars A NS spinning at ∼ ms period has a spin energy Espin ≈ 2.8 × 1052(Pi/1 ms)−2 erg and the spindown timescale through magnetic dipole radiation and/or relativistic par- ticle winds is extremely short, from a few weeks to ∼ one day for dipole fields in the (1014 ÷1015) G range. In stan- dard magnetar scenarios, most of the initial spin energy is expected to be rapidly transferred to the surrounding supernova ejecta through these spindown mechanisms (Thompson et al 2004). Therefore, present-day SNRs around known magnetar candidates should bear the signature of such a large energy injection. For initial spin periods less than 3 ms, the injected energy would be > 3.5 × 1051 erg, making these remnants significantly more energetic than those surrounding ordinary NSs (≤ 1051 erg s−1). The X-ray spectra of the SNRs surrounding known mag- netar candidates (two APXs and two SGRs) studied by Vink et al. (2006), do not show any evidence that their total energy content differs from that in remnants sur- rounding common NSs (≈ 1051 erg): this result con- strains magnetar parameters at birth. Vink & Kuiper 4.1 Strong GW emission at birth? The general expression for the total energy emitted via GWs is given by: E TOT gw = −Z ∞ Egw dt = −Z 0 ωi Egw ω dω (6) ti GWs from newborn magnetars 5 SN ≃ 1051 ergs. Curves are labelled for different values of the ratio (x) of the Fig. 2 Loci in the Bd vs. Bt plane for which E inj GW to magnetodipole torques. Through Eq. (10), it is seen that fixing this ratio is equivalent to defining the corresponding initial spin. The corresponding locus is thus uniquely determined through the first expression in Eq. (9). Note that, given x, the corresponding curve shifts to the right for decreasing values of E inj SN R. The red curve represents the S/N=8 curve of Fig. (1). In the framework of our model, thus, fast spinning newly formed magnetars that do emit ≤ 1051 ergs of spin-down energy through electromagnetic radiation can be expected to be detectable with LIGO II up to the distance of the Virgo cluster. Stated the other way, objects whose GW signal could be detectable by LIGO II up to the distance of the Virgo Cluster would inject - as found in the case of magnetar candidates in the Galaxy - ≤ 1051 ergs in the expanding shells of the surrounding SNR. We insert Eq. (1) and the expression for the GW lumi- nosity of an elliptically distorted, spinning object into Eq. (6) to obtain the following analytical solution: E TOT gw = I Z ωi 0 ω3 ω2 + A dω = I (cid:20) ω2 2 − A 2 ln(ω2 + A)(cid:21)ωi 0 (7) where A = Kd/Kgw. Since E TOT α of the initial spin energy of the NS, we can write: amounts to a fraction gw (1 − α)ω2 i = A ln(cid:18) ω2 i + A A (cid:19) (8) Finally, defining the ratio of the GW over magnetodipole torque as x ≡ (ω2 i /A): x ≈ 2.25 B 4 t,16.3 d,14 P 2 i,2 B 2 1 − α = ln(1 + x) x (9) The above expressions provide us with two relations be- tween the five parameters (ωi, Bd, Bt, x, α). A third one derives from the fact that the remaining fraction (1 − α) of the initial spin energy is available for being transferred to the ejecta through magnetodipole radiation. By as- suming that all this energy is effectively transferred to the SNR, we get: ω2 i = 2 E inj SN R (1 − α)I = 2x E inj SN R ln(1 + x)I (10) where E inj SN R is the spin-down energy injected in the ex- panding shell, constrained by the results of Vink & Kuiper (2006) to be ≃ 1051 erg. We use the results of Lattimer & Prakash (2001), according to which a 1.4 M⊙ NS has a radius R ≃ 12 km and a moment of inertia I ≈ 0.35M R2 ≃ 1.4×1045 g cm2. Note that, by these numbers, one obtains the fol- lowing relation between the measured P and P of NSs and the corresponding value of the dipolar magnetic field at the magnetic pole (our Bd)1: Bd ≃ 4.4 × 1019 (P P ) 1 2 G (11) As can be seen from Fig. (1), dipole magnetic fields stronger than (5 ÷ 6) × 1014 G rapidly quench the ex- pected S/N ratio of GW signals from newly formed mag- netars. Therefore, we restrict our investigation to polar dipole fields Bd < 6. From Eq. (10) we see that, once x is given, both α and ωi are determined (one can indeed use any of the three as the free parameter). The first ex- pression of Eq. (9) thus provides a relation between Bd and Bt. We have repeated this procedure for five values of the ratio of the GW over the magnetodipole torque (x) and, for each of them, obtained the implied values of ωi and α and a curve Bt vs. Bd. Fig. 2 summarizes our results. Loci in the Bd vs. Bt plane for which fast spinning, ultramagnetized magne- tars are consistent with the energetic constraints derived 1 This is a factor 1.5 less than usually assumed with I = 1045 g cm2 and R = 10 Km. Use of the most up to date parameters is required, given the strong dependence of the two competing torques on the exact value of the magnetic fields. 6 Dall'Osso et al. SN is fixed). Details are given in the caption. by Vink & Kuiper (2006) are drawn for the five chosen values of x or, equivalently, the initial spin period (since Einj In summary, we have first identified a range of initial conditions (spin period, internal and external magnetic field), within which newly formed magnetars can be in- teresting targets for next generation GW detectors. Then we have calculated that, within most of that same re- gion, magnetars should emit less than 1051 ergs through magnetodipole radiation (cfr. Dall'Osso et al. (2007) and Fig. 2). This is compatible with the limits inferred through recent X-ray observations of SNRs around present-day magnetar candidates. 5 Discussion The energy liberated in the 2004 December 27 flare from SGR 1806-20, together with the likely recurrence rate of these events, points to a magnetar internal field strength of ∼ 1016 G or greater. Such a field likely results from dif- ferential rotation in a millisecond spinning proto-magnetar and deforms the star into a prolate shape. Magnetars with these characteristics are expected to be very pow- erful sources of gravitational radiation in the first days to weeks of their life. An evolving periodic GW signal at ∼ 1 kHz, whose frequency halves over weeks, would unambiguously reveal the early days of a fast spinning magnetar. Prospects for revealing their GW signal depend on the birth rate of these objects. The three associations between an AXP and a supernova remnant (ages in the 103 ÷ 104 yr range) implies a magnetar birth rate of ≥ 0.5 × 10−3 yr−1 in the Galaxy (Gaensler et al. 1999). Therefore the chances of witnessing the formation of a magnetar in our Galaxy are slim. A rich cluster like Virgo, containing ∼ 2000 galaxies, is expected to give birth to magnetars at a rate of ≥ 1 yr−1. A fraction of these might have sufficiently high toroidal fields that a detectable GW is produced. It has been recently found that SNR shells around some magnetar candidates have comparable expansion ener- gies to standard SNR shells. This implies that either the NS was not initially spinning as fast as required for an α − Ω dynamo to amplify its field to magnetar strengths, or that most of its initial spin energy was emitted in a way that did not interact with the ejecta. GWs have indeed such property. If the internal magnetic field of newly formed magnetars is > 1016 G, comparable to the lower limit estimated through the enegetics of the Dec 27 Giant Flare, then their GW emission can be strong enough to radiate away most of their initial spin energy. The required amount of energy emitted through GWs is indeed such that, had these magnetar candidates been at the distance of the Virgo cluster, they would have been revealed by a LIGO II-class GW detector. Therefore, GWs from newly formed magnetars can account naturally for the recent X-ray observations of SNRs around galactic magnetars. The main conclusion that can be drawn at present is that newborn, fast spin- ning magnetars represent a potential class of GW emit- ters over Virgo scale distances that might well be within reach for the forthcoming generation of GW detectors. References Abbott,B. et al. [LIGO Scientific Collaboration] gr-qc/0508065 (2005) Alpar, A., & Pines, D. Nature, 314, 334 (1985) Alpar, A., & Sauls, J., A. Ap.J., 327, 723 (1988) Arzoumanian, Z., Cordes, J. M. & Chernoff, D. F. Ap.J., 568, 289 (2002) Bonazzola, S. & Marck, J.A. Ann. Rev. Nucl. Part. Sci., 45, 655 (1994) Bonazzola, S. & Gourgoulhon, E. A.&A., 312, 675 (1996) Brady, P., Greighton, T., Cutler, C. & Schutz, B.F. Phys. Rev. D, 57, 2101 (1998) Brady, P.R. & Creighton, T. Phys. Rev. D , 61, 082001 (2000) Cutler, C. Phys. Rev. D, 66, 084025 (2002) Cutler, C., Golami, I. & Krishnan, B. gr-qc/0505082 (2005) Dall'Osso, S., Re, V., submitted to Phys. Rev. D (2006) Dall'Osso, S., et al., in preparation Duncan, R., C. Ap.J.Letters, 498, L45 (1998) Duncan,R.,C., & Thompson,C. Ap.J.Letters, 392, L9 (1992) Feroci, M., et al. Ap.J.Letters, 515, L9 (1999) Ferrario, L.,& Wickramasinghe, D. M.N.R.A.S., 367, 1323 (2006) Gaensler, B., M., Gotthelf, E., V., & Vasisht, G. Ap.J.Letters, 526, L37 (1999) Hurley, K., et al. Nature, 397, , 41 (1999) Hurley, K., et al. Nature, 434, 1098 (2005) Jaranowski, P., Krolak, A., & Schutz, B.F. Phys. Rev. D, 58, 063001 (1998) Jones, D., I. 2002, Class.Quantum Grav., 19, 1255 (2002) Jones, D. I., & Andersson, N. M.N.R.A.S., 331, 203 (2002) Konno, K., Obata, T., & Kojima, Y. A.&A., 356, 234 (2002) Kouveliotou, C., et al. Nature, 393, 235 (1998) Krishnan, B., Sintes, A.M., Papa, M.A., Schutz, B.F., Frasca, S., & Palomba, C. Phys. Rev. D, 70 082001 (2004) Lattimer, J., M., Prakash, M., Ap.J., 500, 426 (2001) Lyutikov, M. M.N.R.A.S., 346, 540 (2003) Lazzati, D., Ghirlanda, G. & Ghisellini, G. M.N.R.A.S., ,, in press (astro-ph/0504308 v2) (2005) Mazets, E., P., Golentskii, S., V., Ilinskii, V., N., Aptekar, R., L., Guryan & Iu., A. Nature, 282, 587 (1979) Mereghetti, S., & Stella, L. Ap.J.Letters, 442, L17 (1995) Mestel, L., Takhar, H. S. M.N.R.A.S., 156, 419 (1972) Nakar, E., Gal-Yam, A., Piran, T., Fox, D., B., astro-ph/0502148v1 (2005) Owen, B., J., & Lindblom, L. Class.Quantum Grav., 19 1247 (2002) Palomba, C., A.&A., 367, 525 (2001) Popov, S.B. & Stern, B. E. M.N.R.A.S., in press (astro-ph/0503532 v3) (2005) Stella, L., Dall'Osso, S., Israel, G. L. & Vecchio, A. Ap.J.Letters, 634, L165 (2005) Tanvir, N. R., Chapman, R., Levan, A. J. & Priddey, R. S. Nature, 438, 991 (2005) Terasawa, T., et al. astro-ph/0502315 (2005) Thompson, C., & Duncan, R.,C. Ap.J., 408, 194 (1993) GWs from newborn magnetars 7 Thompson, C., & Duncan, R. C. M.N.R.A.S., 275, 255 (1995) Thompson, C., & Duncan, R. C. Ap.J., 473, 322 (1996) Thompson, C., & Duncan, R. C. Ap.J., 561, 980 (2001) Thompson, C., & Murray, N. C. Ap.J., 560, 339 (2001) Terasawa, T., et al. Nature, 434, 1110 astro-ph/0502315 (2005) Thompson, T., A., Chang, P., Quataert, E. Ap.J., 611, 380 (2004) Thorne, K.S. LIGO Internal Document G000025-00-M (2000) L. Ushomirsky, Bildsten, G., Cutler, C., & M.N.R.A.S., 319, 902 (2000) Vink, J., & Kuiper, L. M.N.R.A.S., 370, L14 (2006) Woods, P., M., & Thompson, C. (astro-ph/0406133 v3) (2004)
astro-ph/0103027
1
0103
2001-03-02T00:33:55
An Adaptive Optics Survey for Companions to Stars with Extra-Solar Planets
[ "astro-ph" ]
We have undertaken an adaptive optics (AO) imaging survey of extrasolar planetary systems and stars showing interesting radial velocity trends from high precision radial velocity searches. AO increases the resolution and dynamic range of an image, substantially improving the detectability of faint close companions. This survey is sensitive to objects less luminous than the bottom of the main sequence at separations as close as 1 arcsec. We have detected stellar companions to the planet bearing stars HD 114762 and Tau Boo. We have also detected a companion to the non-planet bearing star 16 Cyg A.
astro-ph
astro-ph
Planetary Systems in the Universe: Observation, Formation and Evolution ASP Conference Series, Vol. 3 × 108, 2000 A.J. Penny, P. Artymowicz, A.-M. Lagrange, and S.S. Russell, eds. An Adaptive Optics Survey for Companions to Stars with Extra-Solar Planets James P. Lloyd, Michael C. Liu, James R. Graham, Melissa Enoch, Paul Kalas, Geoffrey W. Marcy, Debra Fischer Department of Astronomy, University of California, Berkeley, CA 94720, USA Jennifer Patience, Bruce Macintosh, Donald T. Gavel, Scot S. Olivier, Claire E. Max Institute of Geophysics and Planetary Physics, Lawrence Livermore National Laboratory, Livermore, CA 95064, USA Russel White, Andrea M. Ghez, Ian S. McLean Department of Physics and Astronomy, University of California, Los Angeles, CA 90095, USA Abstract. We have undertaken an adaptive optics imaging survey of extrasolar planetary systems and stars showing interesting radial veloc- ity trends from high precision radial velocity searches. Adaptive Optics increases the resolution and dynamic range of an image, substantially improving the detectability of faint close companions. This survey is sen- sitive to objects less luminous than the bottom of the main sequence at separations as close as 1′′. We have detected stellar companions to the planet bearing stars HD 114762 and Tau Boo. We have also detected a companion to the non-planet bearing star 16 Cyg A. 1. Introduction The Lick Adaptive Optics System is a Shack-Hartmann Laser Guide star AO system on the Lick Observatory Shane 3m telescope at Mt Hamilton, California (Max et al. 1997). For these observations, we used the system in natural guide star mode, with the bright star serving as wavefront reference. Under good conditions, the system produces diffraction limited images (0.′′15 FWHM) with a strehl ratio of 0.7 in the K band (2.2 µm). The AO system feeds a 256×256 pixel infrared camera, IRCAL (Lloyd et al. 2000), which reimages the field of view at 0.′′076 per pixel. The camera incorporates a cold focal plane with an occulting finger to obtain high dynamic range images. For this program, we typically take a few minutes of integration of unsaturated images to obtain coverage close to the star, and deep exposures in coronagraphic mode to detect faint companions at larger separations. We have selected targets from those stars with radial velocity planets, or with radial velocity trends from the Lick and Keck radial velocity surveys. 1 2 J. P. Lloyd et al. 2. Results HD 155423 (see Fig 1) shows substantial scatter in precision radial velocity measurements. High resolution imaging shows HD 155423 is a hierarchical triple, with a 0.′′2 close binary (8AU projected separation) M3 dwarf pair, separated 1.′′5 from the F8 dwarf primary. A faint object was previously discovered near 16 Cyg A, but was not known to be physically associated (Hauser & Marcy 1999). We have confirmed by proper motion measurements that the ∆K=5.4 object is a physically associated M5 dwarf (see Fig 1). Radial velocity measurements show a shallow linear trend. HD 114762 has a radial velocity companion with an 84 day period and Msin i=11 MJ up (Latham et al. 1989). We have detected a ∆K=7.3 compan- ion 3.′′3 from the primary (see Fig 1). JHK photometry and follow up Keck AO/NIRSPEC spectroscopy (McLean et al. 2000) reveal the companion to be a late M subdwarf. The physical association of this companion is confirmed by proper motion over a 2 year baseline (Patience et al. 1998). Radial velocity measurements of HR 6623 show a nearly linear trend over 13 years (see Fig 1). This would classify this object as a poorly determined single lined spectroscopic binary. AO imaging resolves the companion, which is an M5 dwarf. Further radial velocity and astrometric observations will allow accurate mass determinations. Tau Boo hosts an Msin i = 3.9 MJ up, 3.3 day period planet, and shows radial velocity residuals (see Fig 1). It has an M2V companion that was discovered in 1849 at 10.′′3. At present it is at 2.′′82, with 0.′′01 per year of orbital motion. Although it has been suggested that there may be additional companions in the system (Wiedemann, Deming, & Bjoraker 2000), we do not detect any additional companions, and attribute the velocity residuals to the stellar companion. Acknowledgments. This research was supported in part by the National Science Foundation under a cooperative agreement with the Center for Adaptive Optics, Agreement No. AST-987678. Work on the Lick adaptive optics system was performed in part under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory, Contract number W-7405-ENG-48 References Cochran, W., Hatzes, A., Butler, R. P. & Marcy, G. W. 1997, ApJ, 483, 457 Hauser, H. M. & Marcy, G. W. 1999, PASP, 111, 321 Latham, D., Stefanik, R., Mazeh, T., et al. 1989, Nature, 339, 38 Lloyd, J. P., Liu, M. C., Macintosh, B. A., Severson, S. A., Deich, W. T. S. & Graham J. R. 2000, in Proc. SPIE 4008, Optical and IR Telescope Instrumen- tation and Detectors, ed. M. Iye & A. F. Moorwood, (Washington: SPIE) Max, C. E., Olivier, S. S., Friedman, H. W., et al., Science, 277, 1649 McLean, I. S. Wilcox, M. K., Becklin, E. E. et al. 2000, ApJ, 533, L45 Patience, J., Ghez, A. M., White, R. J. et al. 1998, BAAS, 193, 9708 Wiedemann, G., Deming, D., Gjoraker, G. 2000, astro-ph/0007216 APS Conf. Ser. Style 3 HD 114762B: NIRSPEC/Keck AO x u F l 600 400 200 0 ) s m / ( y t i c o e V l HR 6623 −200 RMS = 7.26 m/s Lick data 1990 1995 2000 Time (Years) Tau Boo ) 1 − s m ( y t i c o e V l 300 200 100 0 −100 1988 1990 1992 1994 1996 1998 2000 Time (Years) Figure 1. K band images of stars with detected companions; J band spectra of HD 114762B; Radial Velocity data for HR 6623 and Tau Boo (residual)
astro-ph/0012181
1
0012
2000-12-08T13:36:12
Chemical yields from low- and intermediate-mass stars: model predictions and basic observational constraints
[ "astro-ph" ]
In this work we analyse the role of low- and intermediate-mass stars in contributing to the chemical enrichment of the interstellar medium. First we present new sets of stellar yields basing on the results of updated evolutionary calculations, which extend from the ZAMS up to the end of the AGB phase (Girardi et al. 2000; Marigo et al. 1999a). These new yields, that present a significant dependence on metallicity, are then compared to those of other available sets (Renzini & Voli 1981; van de Hoek & Groenewegen 1997). The resulting differences are explained in terms of different model assumptions -- i.e. treatment of convective boundaries, mass loss, dredge-up, hot-bottom burning --, and further discussed on the basis of important empirical constraints which should be reproduced by theory -- i.e. the initial-final mass relation, white dwarf mass distribution, carbon star luminosity function, and chemical abundances of planetary nebulae. We show that present models are able to reproduce such constraints in a satisfactory way.
astro-ph
astro-ph
A&A manuscript no. (will be inserted by hand later) Your thesaurus codes are: 06(08.05.3; 08.16.4; 08.13.2; 09.16.1; 09.01.1; 11.01.1) ASTRONOMY AND ASTROPHYSICS 0 0 0 2 c e D 8 1 v 1 8 1 2 1 0 0 / h p - o r t s a : v i X r a Chemical yields from low- and intermediate-mass stars: model predictions and basic observational constraints Paola Marigo Dipartimento di Astronomia, Universit`a di Padova, Vicolo dell'Osservatorio 2, I-35122 Padova, Italia Received 14 September 2000 / Accepted Abstract. In this work we analyse the role of low- and intermediate-mass stars in contributing to the chemical enrichment of the interstellar medium. First we present new sets of stellar yields basing on the results of updated evolutionary calculations, which extend from the ZAMS up to the end of the AGB phase (Girardi et al. 2000; Marigo et al. 1999a). These new yields, that present a sig- nificant dependence on metallicity, are then compared to those of other available sets (Renzini & Voli 1981; van de Hoek & Groenewegen 1997). The resulting differences are explained in terms of different model assumptions – i.e. treatment of convective boundaries, mass loss, dredge-up, hot-bottom burning – and further discussed on the basis of important empirical constraints which should be repro- duced by theory – i.e. the initial-final mass relation, white dwarf mass distribution, carbon star luminosity function, and chemical abundances of planetary nebulae. We show that present models are able to reproduce such constraints in a satisfactory way. Key words: Stars: evolution – Stars: AGB and post-AGB – Stars: mass-loss – Planetary nebulae: general – ISM: abundances – Galaxies: abundances 1. Introduction Low- and intermediate-mass stars (with masses 0.8 M⊙ <∼ M <∼ 5−8M⊙, depending on model details) play an impor- tant role in galactic chemical evolution, thanks to the ejec- tion into the interstellar medium (ISM) of material con- taining newly synthesized nuclear products, mainly 4He, 12C, and 14N, and possibly 16O. The interest in the nucleosynthetic history of 12C and 14N, in particular, has recently increased thanks to the observations of high red-shift systems. For instance, it is possible to measure the nitrogen abundance and the N/O ratio in the damped Ly-α systems (e.g. Pettini et al. 1995; Lu et al. 1998), whereas carbon is detected in the Ly-α Send offprint requests to: Paola Marigo e-mail: [email protected] forest clouds (e.g. Lu 1991; Crotts et al. 1994, Tytler & Fan 1994). The chemical history of both these elements in a galaxy is quite complex, as it depends, besides other factors, on the relative contributions of stars with different masses, hence releasing their yields at different timescales. In the case of 12C the relative role of low- and intermediate-mass stars, or rather of metal-rich, mass-losing massive stars, in the carbon enrichment at late galactic epochs is still a matter of debate (Prantzos et al. 1994). In the case of 14N, large uncertainties affect the theoretical predictions on both the secondary and primary nucleosynthesis of ni- trogen in low- and intermediate-mass stars, and the pos- sible contribution of primary nitrogen from massive stars (Woosley & Weaver 1982, Maeder 1983; see also the review by Maeder & Meynet 2000). It follows that, in order to trace the history of chemical evolution and of star formation in the Universe – which the observations of high red-shift systems are intended to infer – it is quite important to analyse in detail the nucleosynthesis of these elements taking place in low- and intermediate-mass stars. The contribution of these stars to the chemical enrichment of the ISM essentially occurs during the RGB and AGB phases, both characterized by the occurrence of mass loss and events of surface chemical pollution (dredge-up episodes). Whereas modelling the evolution of a star on the RGB is rather easy by means of modern stellar evolutionary codes, dealing with the most advanced stages – charac- terised by the occurrence of thermal pulses (the TP-AGB phase) – is quite a difficult task, due to both the high complexity of the physics involved, and the remarkable requirement of computing time. Then, an alternative theo- retical approach is offered by synthetic models, which sum- marise the results of complete stellar calculations through simple and practical analytical relations. This allows quick computing of the models and ready analysis of the results. Historically, the first most significant AGB synthetic models were those developed by Iben & Truran (1978) and Renzini & Voli (1981). These pioneer works focused on the contribution of low- and intermediate-mass stars to the element enrichment of the ISM. In particular, the 2 P. Marigo sets of stellar yields presented by Renzini & Voli (1981) have been extensively used in chemical evolution models of galaxies, providing almost the only available data source up to recent years. It is important to remark that in Renzini & Voli (1981) synthetic AGB model the fundamental parameters (essen- tially mass-loss and dredge-up) were specified according to the indications from the current knowledge of the involved physical processes and available complete stellar calcula- tions. In this sense the model was uncalibrated. That ap- proach had the merit of supplying a testing tool for com- plete stellar models, as it pointed out at some fundamen- tal inadequacies in the predictions (e.g. too low efficiency of the third dredge-up), and assumed prescriptions (e.g. too low mass-loss rates), which led to clear discrepancies between theory and observations (see, for instance, Iben 1981; Iben & Renzini 1983; Bragaglia et al. 1995). With awareness of that, the later synthetic AGB mod- els (e.g. Groenewegen & de Jong 1993; Marigo et al. 1996, 1999a) have made the next step ahead, that is update the input prescriptions and calibrate the model parameters in order to reproduce fundamental observables (e.g. the carbon star luminosity functions, the initial-final mass re- lation). Basing on the results of these model calibrations, various sets of stellar yields from low- and intermediate- mass stars have been presented in recent years as an al- ternative to Renzini & Voli (1981), namely: Marigo et al. (1996, 1998), van de Hoek & Groenewegen (1997); see also Forestini & Charbonnel (1997), and Boothroyd & Sack- mann (1999). To the above reference list we will now add the new homogeneous sets of stellar yields presented in this work. With respect to our previous calculations (Marigo et al. 1996, 1998), the present yields are derived from stellar models with updated input prescriptions and improved treatment of the relevant processes involved (i.e. the third dredge-up, see Marigo et al. 1999a for all details). Specifically, we follow the evolution of low- and intermediate-mass stars, coupling the results of complete stellar models (Girardi et al. 2000) – that cover the evolu- tion from the zero age main sequence (ZAMS) up to the onset of the thermally pulsing asymptotic giant branch (TP-AGB) – with synthetic TP-AGB models (Marigo 1998ab; Marigo et al. 1999a) that extend the calculations up to the end of this phase. Relevant model prescriptions are briefly recalled in Sect. 2. With the aid of these evolutionary calculations, we then derive the stellar yields (H, He, and main CNO ele- ments; refer to Sects. 3 and 4) for a dense grid of initial stellar masses (in the range 0.8M⊙ – 5M⊙) and various metallicities (Z =0.004, 0.008, 0.019). For the most mas- sive stars, experiencing hot-bottom burning (hereinafter HBB, or envelope burning) during the TP-AGB phase, the corresponding yields are given for three values of the mixing-length parameter, i.e. α = 1.68, α = 2.00, and α = 2.50. This parameter mainly affects the predicted production of 14N and 4He due to HBB (See Sect. 4.1). The final part of this work (Sect. 5) is dedicated to compare our results with the yields calculated by other au- thors, i.e. Renzini & Voli (1981), and van de Hoek & Groe- newegen (1997). In the attempt to single out the causes of the main differences, we analyse the effect of different model prescriptions (e.g. mass-loss, dredge-up, HBB) on the predicted yields, testing at the same time the capa- bility of a model to satisfy basic observational constraints (e.g. initial-final mass relations, white dwarf mass distri- bution, carbon star luminosity functions, chemical abun- dances of planetary nebulae). 2. Evolutionary models 2.1. Some definitions Let us first define some quantities which will be often used throughout this paper to indicate critical masses for the occurrence of particular physical processes. These quanti- ties are: MHeF, Mup, M min dred, and M min HBB, M min . c The first one, MHeF, denotes the maximum initial mass for a star to develop a degenerate He-core, hence expe- rience the He-flash at the tip of the Red Gian Branch (RGB), and is comprised between 1.7 – 2.5 M⊙ depend- ing on metallicity and model details. The second one, Mup, is defined as the critical stellar mass over which carbon ignition occurs in non-degenerate conditions, marking the boundary between intermediate- mass and massive stars. It is worth recalling that this mass limit is usually comprised within 5 – 8 M⊙, being signif- icantly affected by the adopted treatment of convective boundaries, as discussed in Sect. 5. The third one, M min HBB, corresponds to the minimum initial mass for a star to undergo HBB during the TP- AGB phase. Stellar evolution calculations indicate that Mi >∼ M min HBB ∼ 3.5 − 4.5M⊙, depending on metallicity (see, for instance, Marigo 1998b). The fourth one, M min dred, denotes the minimum initial mass for a star to experience the third dredge-up during the TP-AGB phase. Observations of carbon stars suggest that M min dred ∼ 1.1 − 1.5M⊙, decreasing with the metallicity (see, for instance, Marigo et al. 1999a). c Finally, we recall the basic parameters of the third dredge-up: λ and M min . The parameter λ is intended to measure the efficiency the third dredge-up, being defined as the fraction of the increment of core mass over an inter- pulse period which is dredged-up to the surface at the sub- sequent thermal pulse. The actual values for λ are still a matter of debate among theoreticians, and can range from λ ∼ 0 to λ >∼ 1, depending both on stellar properties (e.g. mass and metallicity) and model details (e.g. treatment of convective boundaries). The parameter M min refers to the minimum core mass for the occurrence of the third dredge-up, and is affected c Yields from low- and intermediate-mass stars 3 by several factors as well. In general, theoretical models would predict that M min decreases – so that the third dredge-up is favoured – at increasing mass and mixing- length parameter, and decreasing metallicity (see, for in- stance, Wood 1981; Marigo et al. 1999a). c 2.2. Input physics In this work we consider low- and intermediate-mass stars, i.e. those with initial masses in the range from about 0.8M⊙ to Mup ∼ 5.0M⊙ and and three choices of the original compositions (i.e. [Y = 0.273, Z = 0.019], [Y = 0.250, Z = 0.008], [Y = 0.240, Z = 0.004]). Their evolution from the ZAMS up to the beginning of the TP-AGB phase is taken from the Padua stellar mod- els (Girardi et al. 2000), that include moderate overshoot from core and external convection (Chiosi et al. 1992; Alongi et al. 1993). The reader should refer to Girardi et al. (2000) for more details of the adopted input physics. The models by Girardi et al. (2000) provide the ex- pected changes in the surface abundance of several chem- ical elements (H, 3He, 4He, 12C, 13C, 14N, 15N, 16O, 17O,18O, 20Ne, 22Ne, 25Mg) caused by the first and second dredge-up episodes, i.e. prior to the onset of the TP-AGB phase. It should be also remarked that these models do not assume any ad-hoc "extra-mixing mechanism", e.g. the so-called cool-bottom process (Wasserburg et al. 1995; see also Charbonnel 1995; Boothroyd & Sackmann 1999; Weiss et al. 2000), which is invoked to reconcile discrepant predictions of surface abundances with those measured in field Population II stars, galactic globular clusters, and Magellanic Clouds clusters (see, for instance, the results by Gratton et al. 2000 in their chemical analysis of field metal-poor stars). Mass loss by stellar winds suffered by low-mass stars (with Mi ≤ MHeF ∼ 1.7−2.2M⊙) on the ascent of RGB, is analytically included applying the classical Reimers (1975) formula to the evolutionary tracks calculated at constant mass by Girardi et al. (2000). An efficiency parameter η = 0.45 is adopted to fulfil the classical observational con- straint provided by the morphology of horizontal branches in Galactic Globular Clusters. Finally, once the the first significant thermal pulse is singled out in each evolutionary sequence calculated by Girardi et al. (2000), that point is assumed to de- fine the starting conditions for synthetic calculations of the TP-AGB phase (following the model prescriptions de- scribed in Marigo 1998ab, and Marigo et al. 1999a), which are carried on up to the complete ejection of the enve- lope by stellar winds. Mass loss is included according to the semi-empirical formalism developed by Vassiliadis & Wood (1993). Nucleosynthesis occurring in the innermost envelope layers of TP-AGB stars with HBB is followed adopting the Caughlan & Fowler (1988) compilation of reaction rates for CNO and p-p reactions. The electron- screening factors are those of Graboske et al. (1973). 3. Derivation of the wind contributions Following the classical definition by Tinsley (1980), the stellar yield, pk(Mi), of a given chemical element k, is the mass fraction of a star with initial mass Mi that is con- verted into the element k and returned to the ISM during its entire lifetime, τ (Mi). Tables A1 – A12 give the quan- tities: My(k) = Mi pk(Mi) (1) expressed in solar masses, for all the chemical elements considered, as function of the initial stellar mass Mi, metallicity Z, and mixing-length parameter α. According to the definition of stellar yield we can write: My(k) = Z τ (Mi) 0 [X(k) − X 0(k)] dM dt dt (2) where dM/dt is the current mass-loss rate; X(k) and X 0(k) refer to the current and initial surface abundance of the element k, respectively. Thanks to the fact that the surface chemical composi- tion of stars not suffering HBB during the TP-AGB phase (i.e. with initial masses Mi <∼ M min HBB), is altered by the oc- currence of discrete and quasi-instantaneous episodes of convective dredge-up, that alternate with periods of con- tinuous mass-loss, the evaluation of stellar yields can be simplified as follows. Denoting by X 1(k), X 2(k), X j(k), the abundance of the species k after the first dredge-up, the second dredge- up, and the jth dredge-up event during the TP-AGB phase, respectively, we get: My(k) = My(k)RGB + My(k)E−AGB + My(k)TP−AGB (3) where My(k)RGB = [X 1(k) − X 0(k)] ∆M ej RGB My(k)E−AGB = [X 2(k) − X 0(k)] ∆M ej E−AGB My(k)TP−AGB = Xj [X j(k) − X 0(k)] ∆M j,ej TP−AGB (4) (5) (6) In Eq. (4) ∆M ej RGB is the mass of the envelope ejected during the entire RGB phase (defined only for low-mass stars). It is worth noticing that most of ∆M ej RGB is lost close to the tip of the RGB, that is after the occurrence of the first dredge-up, so that Eq. (4) is a good approxi- mation of Eq. (2). As far as mass loss on the AGB is concerned, we re- M (Vassil- mark that, with the adopted prescription for iadis & Wood 1993), the amount mass lost during the E-AGB phase is indeed negligible so that we can assume ∆M ej E−AGB = 0 in Eq. (5). The contribution of the TP-AGB phase is evaluated with Eq. (6), that sums all the partial contributions of 4 P. Marigo the pulse cycles, the generic jth one consisting of a ther- mal pulse – when the jth dredge-up possibly occurs – fol- lowed by the inter-pulse period, during which the mass ∆M j,ej TP−AGB is ejected. It should be noticed that this approximation holds for TP-AGB stars which experience only the third dredge-up. For more massive TP-AGB stars (with Mi > M min HBB) also suffering HBB, the changes in the surface chemical com- position and mass loss are concomitant processes, so that the calculation of stellar yields requires the adoption of in- tegration time steps shorter than the inter-pulse periods. In general, negative My(k) correspond to those elemen- tal species which are prevalently destroyed and diluted in the envelope, so that their abundances in the ejected mate- rial are lower with respect to the main sequence values. On the contrary, positive My(k) correspond to those elements which are prevalently produced so that a net enrichment of their abundances in the ejecta is predicted. Figures 1 and 2 show the quantities My(k) for all the chemical elements considered, as a function of Mi and Z. For stars experiencing HBB results are given for three val- ues of the mixing-length parameter. k YCNO = Pk X CNO Finally, for the sake of clarity, we remind that the CNO cycle does not change the total number of CNO nuclei involved as catalysts in the conversion of H into 4He, i.e. /Ak = constant. It follows that the first and second dredge-up, though affecting the surface abundances of the CNO isotopes, do not alter their total abundance by number. In fact, the material injected into the envelope has experienced the CNO cycle involving only isotopes already present in the original composition. On the contrary, the constancy of YCNO breaks down as soon as the dredge-up of primary carbon and oxygen, produced by α-capture reactions at thermal pulses, occurs. However, in all cases the total abundance by mass of the CNO isotopes is somewhat changed because of the conversion of these elements mainly into 14N, so that a small positive CNO yield (in mass fraction), is expected, for example, from the RGB phase (see Tables A1 – A3). The quantity, My(CNO), referring to the total net yield of all CNO isotopes, is shown in Fig. 3. 4. Stellar yields as a function of Mi, Z, and α Model predictions can be understood more easily consid- ering that the stellar yield of a given element is essentially determined by the efficiency and duration/frequency of: – the nucleosynthesis/mixing processes (e.g. dredge-up events, HBB) which alter its abundance in the surface layers; – the mass-loss process which ejects the surface layers into the ISM According to the physical prescriptions adopted in this work for the TP-AGB phase (see also Sect. 5.1) we can summarise the following points: – The third dredge-up determines the surface enrich- ment mainly of 4He, 12C, and 16O. The adopted in- tershell abundances are [X(4He)= 0.76; X(12C)= 0.22; X(4He)= 0.02] according to Boothroyd & Sackmann (1988b). The process is more efficient (i.e. higher λ) at lower Z; – HBB operates via the CNO-cycle, hence essentially in- creasing the surface abundances of 4He, and 14N. The process is more efficient at higher M (provided that M > M min HBB), lower Z, and larger α. – mass loss is, in general, less efficient (i.e. lower M ) at decreasing Z and increasing α. In fact, both factors tend to produce hotter tracks, and generally M anti- correlates with Teff . As a consequence, lower mass-loss rates correspond to longer TP-AGB lifetimes, hence greater number of third dredge-up events and a longer duration of HBB. The expected trend of My(k) for the elements un- der consideration as a function of the stellar initial mass, metallicity, and mixing-length parameter is shown in Figs. 1-3. We can notice the following: – H and 4He The net yields of these elements have mirror-like trends, being negative for H, and positive for 4He. The maximum of 4He production at around 2 − 3M⊙ (depending on metallicity) is explained considering the effect of the increase of the number of thermal pulses (hence dredge-up episodes) with stellar mass for 0.8M⊙ < Mi <∼ 2−3M⊙ (see Fig. 5). This peak is more pronounced at lower metallicities due to longer TP- AGB duration (Fig. 5) and larger number of thermal pulses (Fig. 6) for given stellar mass. The subsequent increase of 4He production towards higher masses, 4M⊙ <∼ Mi ≤ 5M⊙, is caused by the occurrence of HBB in addition to the third dredge-up. The yield of 4He is larger for lower metallicities and higher values of the mixing-length parameter, reflecting the greater efficiency of both the third dredge-up and HBB. – 3He The net yield of this element presents a pronounced peak at very low masses, say at Mi ∼ 1M⊙, and de- creases at higher masses. This trend is explained con- sidering that the main contribution to the yield is due to the first dredge-up, and that both the related surface enrichment of 3He and the amount of mass lost during the RGB phase are inversely proportional to the stel- lar mass in the low-mass domain. At higher masses, (Mi >∼ 3M⊙), the net yield becomes even slightly neg- ative because of HBB. – 12C The net positive yield increases with the initial mass, up to a maximum located at about 2 − 3M⊙, corre- sponding to largest number of dredge-up episodes suf- fered during the TP-AGB phase, provided that HBB has not operated. Yields from low- and intermediate-mass stars 5 Fig. 1. Net yields My(k) (in M⊙) for each indicated chemical element (H, 3He, 4He, 12C, and 13C) as a function of the initial mass (in M⊙) of the star. The solid, dashed, and dotted lines correspond to the metallicity sets Z = 0.019, Z = 0.008, and Z = 0.004, respectively. Panels along each column refer to the same value of the mixing-length parameter. The subsequent decline towards higher masses is ini- tially due to fewer dredge-up events, and then to the prevailing effect of HBB. It follows that no substantial enrichment of 12C is provided from the most massive AGB stars. This general trend is more marked at lower metallic- ities, because of the longer TP-AGB phases, and the greater efficiency of the third dredge-up and HBB. – 13C The first dredge-up causes an increase of the 13C sur- face abundance in stars of all masses, and the resulting yields are positive. A further contribution is provided 6 P. Marigo Fig. 2. Net stellar yields My(k) (in M⊙) for each indicated chemical element (14N, 15N, 16O, 17O, and 18O) as a function of the initial stellar mass (in M⊙). The notation is the same as in Fig. 1. by mild HBB, as long as creation of 13C via the re- action 12C(p, γ)13C prevails over destruction via the reaction 13C(p, γ)14N. The results also depend on the interplay between the strength of HBB and mass loss. The most favourable cases correspond to the mod- els [e.g. (4M⊙, Z = 0.019, α = 2.00), (3.5M⊙, Z = 0.004, α = 1.68)], in which the efficiency of reactions allows the synthesis of 13C for a long time before the drastic reduction of the envelope causes the extinc- tion of nuclear burning. The spikes of 13C production for these models would suggest that a proper tuning of HBB is required: If nuclear reactions are somewhat too weak 13C is not significantly created, else if somewhat too strong 13C is quickly destroyed in favour of 14N. – 14N It turns out that HBB plays the dominant role for the synthesis of 14N. The positive yield as a function of the stellar mass depends on both the efficiency and dura- Yields from low- and intermediate-mass stars 7 Fig. 3. Net stellar yields of all CNO elements as a function of the initial stellar mass. The notation is the same as in Fig. 1. tion of nuclear burning. It follows that lower metallic- ities and higher values of the mixing-length parame- ter concur to favour nitrogen production. The contri- bution from low- and intermediate-mass stars to the galactic enrichment of nitrogen may result important, as suggested by chemical evolutionary models of galax- ies (e.g. Portinari et al. 1998). – 15N The net yield of this element is mostly negative. For stars with initial masses in the range, 0.8M⊙ <∼ Mi <∼ 3.5M⊙, the depletion of 15N is due to the effect of the first and second dredge-up. For higher mass stars, the results are affected by HBB, depending on the degree of CNO cycling attained in the burning regions. This element has the shortest nuclear lifetime, after that of 18O, so that it quickly attains nuclear equilibrium with 14N. We note that the depletion of 15N is much more ef- ficient at higher metallicities due to the increase of the CNO-cycling, which implies a more efficient destruc- tion of this element. The predicted trend of the 15N yield mirrors, to some extent, that of the secondary component of 14N (cf. Sect. 4.1). – 16O The net yield of this element is positive for stars with 0.8M⊙ <∼ Mi <∼ 3.5M⊙ thanks to the dredge-up events during the TP-AGB phase, whereas it becomes nega- tive at higher masses because of HBB. The synthesis of fresh 16O occurs via the reaction 12C(α, γ)16O during thermal pulses, so that the yield of this element depends, among other factors, on its abundance in the dredged-up material. According to recent TP-AGB calculations – which include deep overshooting from all convective boundaries – Herwig et al. (1997) find that the 16O abundance in the con- vective intershell is roughly ten times higher, X(16O)∼ 0.25, than previously predicted, i.e. X(12C)∼ 0.02 by Boothroyd & Sackmann 1988b (standard case). Then, adopting Herwig et al. indications, oxygen production by low- and intermediate-mass stars may be favoured with respect to the standard case, but it is worth re- calling that, in general, the final yields are crucially affected by various other factors (e.g. number and ef- ficiency of dredge-up episodes; see also Sect. 6). Anyhow, regardless of its abundance in the dredged- up material, the trend of 16O, as a function of M and Z, is expected to be qualitatively similar to that of 12C. Moreover, according to the present calculations, a notable dependence on metallicity comes out. The increasing positive trend with decreasing metallicities essentially reflects the longer duration of the TP-AGB phase, and hence the greater number of dredge-up episodes, in combination with their larger efficiency. – 17O A certain production of this element is provided by TP- AGB stars with HBB, as long as the chain of reactions 16O(p, γ)17F(β+ν)17O prevails over nuclear destruc- tion via the reactions 17O(p, γ)18F and 17O(p, α)14N. The behaviour resembles that of 13C, in the sense that under some fine-tuned conditions – for particular com- binations of the stellar mass and metallicity – a pro- duction spike of 17O can result. – 18O This element has the shortest nuclear lifetime against proton captures, so that it easily burns even at mild temperatures, quickly attaining the nuclear equilib- rium with 17O. The net yield of 18O is negative for all masses, with a trend mirroring that of 17O. More- over, as in the case of 15N, the depletion of 18O is more pronounced at higher metallicities due to the more ef- ficient CNO cycling. 8 P. Marigo 4.1. Secondary and primary components As far as the CNO nuclei are concerned, we can distinguish for each element k the secondary, M S y (k), and primary, M P y (k), components of the stellar yield My(k) = M S y (k) + M P y (k) calculated with (see Eq. (2)) M S y (k) = Z τ (Mi) 0 [X S(k) − X S,0(k)] and M P y (k) = Z τ (Mi) 0 [X P(k) − X P,0(k)] (7) (8) (9) dM dt dt dM dt dt We notice that M P y (k) can be only ≥ 0, whereas M S where X S(k) and X P(k) denote the secondary and pri- mary current surface abundances, respectively. Moreover, we have X S,0(k) = X 0(k) and X P,0(k) = 0, as follows from the definitions of secondary and primary abundances. y (k) can be either ≥ 0 or < 0, in the respective cases that the mass-averaged secondary abundance of the element in the ejecta is greater, equal or smaller than its original value. A few basic remarks should be made at this point. Both the first and second dredge-up affect (by increasing or decreasing) only the secondary components of the CNO surface abundances. In fact, in these episodes the envelope is polluted by material which has undergone CNO-cycling, with a net change in the relative abundances of the CNO isotopes synthesized from metal seeds originally present in the star. On the contrary, the third dredge-up enriches the chemical composition of the envelope with 12C and 16O of primary origin (synthesized by α-capture reactions). Fi- nally, HBB affects the abundance distribution of the CNO isotopes of both secondary and primary synthesis. Keeping in mind these concepts, it turns out that: – stars with initial masses Mi < M min dred can produce CNO yields of secondary origin only; – for stars with initial masses M min dred <∼ Mi <∼ M min HBB the yields of 12C and 16O should include a primary component, as a consequence of the third dredge-up during the TP-AGB phase; and – for stars with initial masses Mi >∼ M min HBB undergoing HBB during the TP-AGB phase, the yields of the CNO isotopes should all display primary components, be- cause of the injection of primary 12C and 16O nuclei (at each dredge-up event) into regions where the CNO cycle is operating. Comparing the secondary and primary components of the CNO yields four cases can be met (Tables A1 – A12; see also Marigo 1998b): 1. M S y 6= 0 and M P y = 0 so that My = M S y 2. M S 3. M S 4. M S y > 0 and M P y < 0 and M P y < 0 and M P y > 0 so that My > 0 y > 0 so that My > 0 y > 0 so that My < 0 The first (1.) case applies to low-mass stars with Mi < M min dred, i.e. never experiencing the third dredge-up dur- ing the TP-AGB phase. No primary component of stellar yields is expected. The second (2.) case corresponds to both primary and secondary production. In stars with Mi > M min HBB it ap- plies, for instance, to 13C, 17O, and 14N. In general, for these elements a positive secondary contribution may be provided by the first (and possibly second) dredge-up and HBB, the latter process being also responsible for the pri- mary synthesis of these elements (starting from primary 12C and 16O injected by the third dredge-up). As far as 13C is concerned (see also Sect. 4) we notice that a suitable interplay between the strength of HBB and mass loss can occasionally result in very favourable conditions for the production of 13C, giving a peak of the related yields (for instance, at the model 4M⊙ in the case Z = 0.019, α = 2.0). Concerning the yields of 14N, it should be remarked that the contribution from intermediate-mass stars may be relevant in view of interpreting, with the aid of chem- ical evolutionary models of galaxies, the observed trend in the log(N/O) vs. log(O/H) diagram (see, for instance, Vila-Costas & Edmunds 1993; Henry et al. 2000b), where the large scatter of data points towards lower metallicities would imply the existence of a significant primary compo- nent in the measured nitrogen abundances. The third (3.) possibility corresponds to a dominant <∼ primary production. In stars with initial masses M min dred Mi <∼ M min HBB, this case applies to 12C surface abundance, which is first decreased by the negative secondary con- tribution from the first and second dredge-up, and subse- quently increased by the third dredge-up injecting primary nuclei into the envelope. A similar situation occurs for the yield 16O in the same range of stellar masses, as the effect of the third dredge-up prevails over that of the previous mixing episodes (first and second). Finally, the fourth (4.) case corresponds to a dominant secondary depletion. This refers to 15N, 18O for stars of all masses, and to 16O for stars with Mi >∼ M min HBB if the re- duction of the original abundance caused by the first and second dredge-up dominates over the injection of primary oxygen via the third dredge-up (even possibly partially destroyed by HBB). Under these circumstances, no en- richment of the interstellar medium is expected for these elemental species. Tables A10 – A12 give the net yield for each element of the CNO group (T entry), together with the secondary (S entry) and primary (P entry) components, for stars with initial masses 3.5M⊙ ≤ Mi ≤ 5M⊙, for various values of the original metallicity and mixing length parameter. Yields from low- and intermediate-mass stars 9 Table 1. Summary of the main prescriptions adopted in the synthetic TP-AGB models here considered for comparison. PROCESS/QUANTITY MODEL PRESCRIPTION Mup RV81 8 M⊙ HG97 7 M⊙ M2K (this work) 5 M⊙ Mc − L relation, Mc − Tip relation no metallicity dependence with metallicity dependence with metallicity dependence mass loss 3 D.up: efficiency λ 3 D.up: minimum core mass M min c HBB: overluminosity HBB: nucleosynthesis Reimers (1975) η = 1/3 − 2/3 Reimers (1975) η = 5 Vassiliadis & Wood (1993) ∼ 0.3 − 0.5 function of Mc, any Z 0.60 M⊙ for any Z, M no 0.75 for any Z, M 0.58 M⊙ for any Z,M no 0.65 for Z = 0.004, any M 0.55 for Z = 0.008, any M from envelope integrations function of M and Z yes nuclear network parameterised approx. nuclear network 5. Comparison with other calculations We will compare the stellar yields presented in this work (hereinafter also M2K) with those available in two widely used studies, namely the pioneer work by Renzini & Voli (1981; RV81), and the more recent one by van de Hoek & Groenewegen (1997; HG97). Before making a direct com- parison between the yields of various elemental species, we consider it useful first to recall the relevant prescriptions adopted in the mentioned AGB models, and consequently analyse their effects by showing how the predictions of different models compare with basic observables. 5.1. Model prescriptions Table 1 summarises the relevant assumptions adopted in the AGB calculations, which the three different sets of stellar yields under consideration are derived from. For further details the reader should refer to the original pa- pers and references therein. that models with masses up to Mup cover, by definition, the whole class of low- and intermediate-mass stars. In other words, models with Mi > Mup (for whatever pre- dicted Mup) would eventually meet the fate of supernova explosion. Finally, a cautionary warning should be made in the context of practical application of chemical yields. When stellar yields from stars of different initial masses are to be included in galactic models of chemical evolution, at- tention should be paid to correctly match sets of yields of different mass intervals (i.e. low, intermediate, high). If the models do not belong to a homogeneous grid of stellar calculations, one should at least care to combine stellar yields of different origin (i.e. stellar code) but with the same predicted value for Mup. Otherwise, the relative weight of stars belonging to different classes (i.e. with dif- ferent nucleosynthetic histories) to the integrated chemical enrichment may be substantially mistaken (for instance, by over- or under- estimating the role of supernovae). 5.1.1. The limiting mass Mup 5.1.2. Analytical relations First of all, let us consider the quantity Mup (see also Sect. 2.1), that corresponds to the maximum initial mass of a star to develop a degenerate C-O core, hence experi- ence the AGB phase. This critical mass heavily depends on the previous evolutionary history, mainly the exten- sion of the convective core during the H-burning phase. Evolutionary models (classical models) that adopt the Schwarzschild criterion to define the convective bound- aries (e.g. Dominguez et al. 1999) predict higher values for Mup than those models that assume some overshoot beyond the last formally stable layer against convection (e.g. Girardi et al. 2000). Both RV81 and HG97 yields are based on classical models and have Mup ∼ 7 − 8M⊙, whereas the pre-AGB evolutionary models used in this work (M2K) adopt a con- vective overshoot scheme so that Mup ∼ 5M⊙. We remark Fundamental relations in synthetic AGB models are the core mass - luminosity (Mc−L), and core mass - interpulse period (Mc −Tip) relations. Predictions of stellar yields are significantly influenced by these input prescriptions, i.e. the luminosity affects the mass-loss rates on the TP-AGB, and interpulse-periods determine the temporal recurrence of the third dredge-up. With respect to RV81 prescriptions (Iben & Truran 1978 (IT78), see Fig. 4), M2K and HG97 models are based on more recent relations (Boothroyd & Sackmann 1988a; Wagenhuber & Groenewegen 1998) in which a notable im- provement is the inclusion of a composition dependence. For a given core mass, the quiescent luminosity / inter- pulse period of a TP-AGB star is found to increase / de- crease at increasing metallicity. According to evolutionary calculations by Boothroyd & Sackmann (1988a), for in- 10 P. Marigo Fig. 4. Theoretical core mass - luminosity relations. The well-known Iben & Truran (1988) relation is extrapolated for small core masses and assuming a total mass of 1.5 M⊙. The predicted effect of the chemical composition of the en- velope on the luminosity is shown according to Boothroyd & Sackmann (1988a). Fig. 5. Expected number of thermal pulses (and inter- pulse periods) as a function of the initial stellar mass for different values of the initial metallicity (see legenda in Fig. 6). Results of the present work (M2K) are compared with those of Renzini & Voli (1981; RV81). stance, the quiescent TP-AGB luminosity for Z = 0.02 is about 20 % higher than for Z = 0.001 (see Fig. 4). 5.1.3. Mass loss on the AGB The adopted prescription for mass loss on the AGB cru- cially influences the predictions of stellar yields, as it af- fects the total number of thermal pulses (hence dredge-up episodes) suffered by a TP-AGB star, hence the growth of its core mass and AGB lifetime. In this work we adopt the semi-empirical formalism presented by Vassiliadis & Wood (1993) who couple results of pulsation theory with observations of variable AGB stars. In RV81 the classical Reimers' law (1975) is assumed with the efficiency parameter η set equal to 1/3 or 2/3. HG97 as well use the Reimers' law, but with η = 5, which they find as the best value to fulfil basic observational constraints (see Sects. 5.2.2 and 5.2.1). To this regard the following remark should be made. As a matter of fact, the Reimers' prescription was origi- nally designed to describe mass loss suffered by low-mass stars climbing up the RGB, and it is usually calibrated in view of reproducing observations of stars in the subsequent horizontal branch phase of quiescent core He-burning (see Sect. 2). lution does not suit important constraints. In fact, the Reimers' prescription with η <∼ 1 cannot produce the "super-wind" mass-loss rates ( M ∼ 10−4 − 10−5 M⊙ yr−1) measured in stars close to the AGB-tip luminosities, and consequently it cannot account for the typical values of masses and radii of planetary nebulae at the observed luminosities. These difficulties have been overcome by later mass- loss prescriptions – specifically designed for AGB stars (e.g. Bowen 1988; Fleisher et al. 1992; Vassiliadis & Wood 1993; Blocker 1995) – which are all characterised by a more rapid increase of mass-loss rates during the AGB evolution, then naturally leading to the development of the superwind regime. The effect of different laws for mass loss is significant with respect to the expected number of thermal pulses experienced on the TP-AGB evolution. As an example (see Fig. 5), we can notice that for a (5M⊙, Z = 0.02) model our calculations yield Np = 117 − 153 (depending on α), whereas RV81 predict Np = 8941 (1631) with the efficiency parameter η = 1/3 (η = 2/3). In general, it turns out that the largest differences in the number of thermal pulses show up for models with higher stellar masses (i.e. Mi >∼ 2.5M⊙), that are expected to experience the super- wind regime according to the the Vassiliadis & Wood's prescription. However, as already shown by RV81, the straightfor- ward extension of the Reimers' formula to the AGB evo- Moreover, significantly different results are obtained by RV81 and M2K as far as the TP-AGB duration is Yields from low- and intermediate-mass stars 11 cumstance is always prevented by the earlier removal of the whole stellar envelope. 5.1.4. The treatment of the third dredge-up This process is expected to critically affect the actual yields of many elements, mainly 4He, 12C, and 16O. More- over, the reduction of the core mass caused by the third dredge-up concurs to determine the final mass of the rem- nant left at the end of the AGB. The largest differences in the analytical treatment of this process, in the models here considered, can be sum- marised as follows. In RV81 model the onset of the third dredge-up in low-mass stars possibly occurs later (higher M min ) and with a lower efficiency (lower λ) than in the HG97 and M2K calculations, that are carried out with similar values of the parameters. (see Table 1). c c This can be explained considering the different usage of the quantities M min , and λ in the models. As already men- tioned in Sect. 1, in RV81 these parameters were derived according to complete stellar models currently available at that epoch. The subsequent comparison between the pre- dictions of synthetic models and observations pointed out at the so-called "carbon star mystery", as denominated by Iben (1981), i.e. too few faint and too many bright carbon stars expected than observed. c Differently, the later analyses carried out by HG97 and M2K move from another perspective, that is to consider M min and λ as free parameters which should be calibrated in order to reproduce observations of carbon stars. The aim is to provide indications on the average characteris- tics of the third dredge-up, so as to remove the theoreti- cal discrepancy related to the "carbon star mystery" (see Sect. 5.2.3). 5.1.5. The treatment of hot bottom burning The most notable effect of this process on stellar yields involves 14N and 4He, which are newly synthesised at the expense of hydrogen and, in general, of the other CNO catalysts. It is important to stress that the occurrence of HBB in the most massive AGB stars (M > 3.5 − 4.5M⊙) has not only a direct effect on stellar yields – via changing the chemical abundances in the envelope – but also deter- mines an indirect action, affecting the energetics of these stars, hence their evolutionary properties. In fact, as a consequence of HBB, the Mc − L relation breaks down, a feature clearly shown by complete AGB stellar calculations (e.g. Blocker & Sconberner 1991; see also Fig. 7). In these massive AGB models the luminosity evolution is characterised by a steeper increase with the core mass (above the Mc − L relation) up to a maximum, followed by a decline as soon as the envelope mass is sig- nificantly reduced by mass loss. Eventually the Mc − L re- lation is recovered (e.g. Vassiliadis & Wood 1993; Marigo 1998b). This behaviour is exemplified in Fig. 7 where we Fig. 6. Predicted TP-AGB lifetimes as a function of the initial stellar mass and metallicity according to the present work (M2K) and that of Renzini & Voli (1981; RV81) with the mass-loss parameter η = 1/3. concerned (see Fig. 6). First of all, we can notice that our TP-AGB lifetimes present a pronounced trend with the stellar mass, showing a maximum at around 2 – 2.5 M⊙ (depending on metallicity). In particular, a drastic drop of the TP-AGB duration is expected in the high- est mass domain, i.e. for stars with HBB. In RV81 the mass-dependence is much less marked, and such a strong reduction of the TP-AGB lifetimes of the most massive models is not predicted. This latter point is relevant for the interpretation of the high-luminosity wing of the ob- served carbon star luminosity functions (see Sect. 5.2.3). We also notice that with the Vassiliadis & Wood's for- malism both the total number of the thermal pulses and the TP-AGB lifetimes are notably sensitive to the metal- licity, i.e. increase with decreasing Z. This feature reflects consequently on the predicted yields from stars with the same initial masses but different initial metallicities. Finally, it is worth remarking that the efficiency of mass loss on the AGB crucially affects the masses of the bare C-O cores left at the end of this phase. It follows that the empirical initial-final mass relation sets impor- tant constraints to the theoretical prescriptions for stellar winds (see Sect. 5.2.1). It is important also to recall that according to RV81 in stars with Mi >∼ (5 − 8)M⊙ the C- O degenerate core grows up to the Chandrasekhar limit (∼ 1.4M⊙; see Fig. 8), leading to explosive carbon ignition (Type I-1/2 Supernova event; see Iben & Renzini 1983). On the contrary, according to M2K and HG97 this cir- 12 P. Marigo report the results of complete evolutionary calculations performed by Blocker (1995) for a 7M⊙ model with solar metallicity (triangles). Actually, the effect of the predicted luminosity evolu- tion on stellar yields is at least two-fold. In fact, the stellar luminosity is closely related to the temperature at the base of the envelope, which the nuclear reaction rates crucially depend on. Moreover, the overluminosity of AGB stars with HBB can trigger high mass-loss rates, thus favour- ing the onset of the super-wind regime with consequent reduction of the TP-AGB lifetimes. Such overluminosity effect caused by HBB is included neither in RV81 nor in GdJ93, where the luminosity evo- lution is assumed to follow the Mc − L relation by IT78 (with some revision for the composition dependence in the GdJ93 work). In Fig. 7 we show the behaviour of the luminosity for the 7M⊙ model (solid line) as it would re- sult adopting the IT78 relation with the the same values for current Mc and M as in the Blocker (1995) model se- quence. The discrepancy is notable. To overcome this lim- itation of synthetic models Marigo et al. (1998; see also Marigo 1998ab) developed a solution scheme based on en- velope integrations, so that the overluminosity produced by HBB is taken into account and the results of complete stellar calculations are recovered (dashed line in Fig. 7). 5.2. Observational constraints In the following we will examine the AGB synthetic mod- els under consideration (RV81, HG97, M2K (this work)) in relation to their capability of reproducing important observational constraints, which are closely related to the stellar yields. 5.2.1. The initial - final mass relation The initial-final mass relation (IFMR) of low- and intermediate-mass stars is intimately linked to the chem- ical yields, as it determines the total amount of matter ejected by a star during its entire evolution. In a com- plementary way, it gives information on the reservoir of stellar remnants, irreversibly lost by the star-forming gas. Moreover, assessing the upper mass limit for WD progen- itors (MWD) is an important point, since it affects the expected rate of type II supernovae. All these aspects are fundamental issues for chemical evolutionary models. Figure 8 shows a few empirical calibrations of the IFMR for the solar neighbourhood. The first striking point is that the more recent determinations significantly differ from the earlier work by Weidemann (1987). For instance, the revised relation by Herwig (1996) presents a flatter slope up to Hyades location (Mi ∼ 3M⊙, Mf ∼ 0.7M⊙), followed by a steeper rise, and a final flattening towards higher initial masses (Mi > 4M⊙). The presence of an inflection point at the Hyades mean location seems to be Fig. 7. Quiescent luminosity as a function of the core mass during the TP-AGB phase. The predictions for the 7 M⊙, Z = 0.021 model with HBB according to full evo- lutionary calculations by Blocker (1995) are compared to the those of synthetic calculations. The luminosity evolu- tion for a 2.5M⊙ model is also shown (Marigo et al 1999a). The reference Mc − L relation is taken by Wagenhuber & Groenewegen (1998). The adopted mass-loss prescription is that by Baud & Habing (1983). See text for further explanation. confirmed also by Reid (1996), as discussed by Weidemann (1997). The second point to be made deals with the critical 1, that is the maximum initial mass of WD mass MWD progenitors. At present, this limiting value is still rather uncertain (most likely in the range 5 – 8 M⊙), since it heavily depends on model details. In particular, as already discussed by Weidemann (1987), the definition of convec- tive boundaries – via either the Schwarzschild criterion or an overshooting scheme – plays a crucial role. It turns out that with the latter choice MWD is lower than assuming the former classical assumption. However, other param- eters may affect the predictions for MWD. For instance, the recent metallicity re-determination (i.e. half-solar) of the young open cluster NGC 2516 by Jeffries (1997; see Fig. 8) has lead to assign it a younger age. As a conse- quence, Jeffries (1997) derives MWD around 5 – 6 M⊙, that is considerably lower than MWD ∼ 7 − 8 M⊙ as esti- 1 For the sake fo simplicity, we limit our discussion to MWD for carbon/oxygen white dwarfs, i.e. not considering the neon/oxygen white dwarfs which would derive from "super- AGB" stars according to Garcia-Berro et al. (1997) evolution- ary calculations. Yields from low- and intermediate-mass stars 13 Fig. 9. WD mass distribution for solar metallicity. The observed data in the solar neighbourhood are taken from the sample of bright DA WDs by Bragaglia et al (1995). The shaded area beneath the ob- served histogram is set equal to unity. The middle and right pan- els on the top show the WD mass distributions (solid line) derived by adopting the semi-empirical IFMRs by Weidemann (1987) and Herwig (1996), respectively. The bottom panels display the distributions ob- tained by assuming the theoretical IFMRs from the quoted works. All the solid line histograms are nor- malised to the fraction of observed WDs with masses larger than 0.5 M⊙. See text for more details. more recent paper Jeffries et al. (1998) still do not exclude that the metallicity of NGC 2516 might be nearly solar. Figure 8 displays the theoretical IMFRs as derived by RV81, HG97, and M2K for low- and intermediate mass- models with initial solar metallicity. We can notice that the both HG97 and M2K are satisfactorily consistent with the trend of the most recent observational relations, whereas RV81 is far from reproducing the empirical data. In particular, the RV81 relation shows a quick divergency of the final mass at increasing initial mass, with the most massive stars being able to build C-O cores up to the Chandrasekhar limit of 1.4 M⊙. The final fate of these stars would correspond to the occurrence of type I-1/2 su- pernova events, which seems not to be supported by the observations. 5.2.2. The white dwarf mass distribution The white dwarf mass distribution (WDMD) is closely related to the IFMR from which it can be theoretically derived, provided that assumptions on the initial mass function (IMF) and star formation rate (SFR) are made to perform the integration over mass and time. The observed WDMD in the solar neighbourhood (Bergeron et al. 1992; see Fig. 9) is narrowly peaked in the mass range 0.5 – 0.6 M⊙ (adopting a mass bin of 0.1 M⊙), which contains more than 45 % of the observed objects. We notice that, with a finer bin sampling (i.e. 0.05 M⊙), the location of observed peak would fall between 0.5 and 0.55 M⊙, which may be difficult to be theoreti- cally reproduced. In fact, according to stellar evolutionary models, these values would be consistent with the mini- mum remnant mass produced by progenitor stars as old as the age of the Galaxy (∼ 15 Gyr corresponding to ini- relation for low- Initial-final mass Fig. 8. and intermediate-mass stars with solar metallicity. Semi- empirical calibrations for the solar neighbourhood are taken from Weidemann (1987), Herwig (1996), and Jeffries (1997). Solid lines refer to theoretical predictions. When the Reimers' prescription for mass-loss is adopted, the corresponding efficiency parameter η is indicated. See text for more details. mated in previous studies (e.g. Weidemann 1987; Koester & Reimers 1996). However, it should be noticed that in a 14 P. Marigo The predicted fraction of WDs contained in the kth mass bin (from M k f to M k+1 f ) is calculated with: N k ∝ i (M k+1 (M k+1 − M k i ) − M k f ) f i , M k+1 i φ(M k/2 i ) ψ(TG − τk/2) ∆Tk/2 (10) i where (M k ) is the corresponding interval of initial stellar mass; φ(M k/2 ) is the IMF (by number) evaluated at the mean initial mass 0.5 (M k+1 i ); TG and τk/2 denote the age of the Galaxy, and the lifetime of a star with mass M k/2 , respectively; ψ(TG − τk/2) is the SFR evaluated at the birth epoch of the star; and ∆Tk/2 is the time interval of detectability of the WD. + M k i i For the sake of simplicity, in our calculations we adopt a constant SFR, the IMF given by the Salpeter's law (φ(M ) ∝ M −2.5), and suppose that any WD formed in the past is still detectable at the present time. This im- plies we assume that the WD fading time is always much longer than the WD's age. In this case ∆Tk/2 = TG − τk/2, i.e. the WD has been detectable since the death of the pro- genitor. Moreover, it is worth noticing the following points. Un- der the assumption of a constant SFR, the WDMD essen- tially depends on i) the slope, dMi/dMf, of the IFMR, ii) the IMF, and iii) the lifetimes of the stellar progenitors relative to the age of the Galaxy. The first factor favours the population of those mass bins in which the slope of the IFMR is flatter, i.e. where stars with different initial masses build up WDs with similar masses. This should be one of the dominant ef- fects which gives rise to the observed narrow peak of the WDMD at around 0.5 − 0.6 M⊙, just where the IFMR is rather flat (see Fig. 8). The second and third factors tend to produce opposite effects. The IMF preferentially weighs the formation of WDs with lower masses, hence evolved from originally less massive stars if the IFMR is a monotonic function. On the contrary, the accumulation factor (TG − τ ) favours the contribution of WDs of higher masses, evolved from more massive stars, hence with shorter lifetimes. In Fig. 9 the observed WDMD in the solar neighbour- hood is compared with the distributions derived accord- ing to Eq. (10) assuming different IFMRs, both empir- ical and theoretical ones. The relations by Weidemann (1987) and Herwig (1996), though being quite different, yields WDMDs both acceptably consistent with the ob- served one. This can be explained considering that the major differences between the two IFMRs show up for Mi > 2.5 − 3M⊙, corresponding to WDs that do not con- tribute to the mass peak. For Mf ≤ 0.55M⊙ the relations are quite similar, showing a rather flat trend. As far as the purely theoretical WDMDs are concerned (bottom panels), it turns out that a satisfactory reproduc- tion of the observed data is attained by both HG97 and M2K (this work), whereas a notable discrepancy affects the predictions by RV81. As already anticipated in the Fig. 10. Luminosity functions of field carbon stars in the Magellanic Clouds. The observed data (shaded his- tograms) are taken from Costa & Frogel (1996) for the LMC, and Reiberot (1993) for the SMC. Theoretical dis- tributions (thick solid line) are shown for comparison. Top left-hand side panel: Iben & Renzini (1983) calculations for the LMC – based on a synthetic AGB model very sim- ilar to that of Renzini & Voli (1981; RV81) – with the pa- rameter set (η = 2/3, α = 2, ǫ = 0.1). Top right-hand side panel: Groenewegen & de Jong (1993; GdJ93) best fitting distribution for the LMC carbon stars. Bottom panels: Marigo et al. (1999a) best fits to the CSLFs in the LMC and SMC. See text for further explanation. tial masses ∼ 0.9 M⊙), provided that their C-O cores do not grow in mass during the TP-AGB phase. In fact, these stars are predicted to enter the TP-AGB phase with a core mass of already ∼ 0.52 M⊙ (Girardi et al. 2000). Then, the location of the observed peak in the range 0.5 – 0.55 M⊙ might be explained by theory only if assuming i) that mass loss suffered by AGB low-mass stars is so strong that they leave this phase as soon as they enter it, or ii) very effi- cient dredge-up prevents the growth in mass of the core (Herwig et al. 1997). On the other hand, as suggested by Bragaglia et al. (1995), the origin of such discrepancy be- tween theory and observations would be most likely due to a systematic underestimation of the surface gravities derived from WD models. Given this point of uncertainty and considering that WDs with M < 0.4M⊙ are probably helium WDs de- rived from binary evolution, in this work both theoretical and observed WDMDs (see Fig. 9) are derived adopting a mass bin of 0.1 M⊙, and normalising them to the observed fraction of WDs with M ≥ 0.5M⊙. Yields from low- and intermediate-mass stars 15 discussion on the predicted IFMRs (Sect. 5.2.1), in RV81 there is a sizable overproduction of WDs more massive than 0.6M⊙, a feature already pointed out by Bragaglia et al. (1995). 5.2.3. The carbon star luminosity function The carbon star luminosity function (CSLF) is a funda- mental observable as it gives indications on, at least, two basic processes occurring in TP-AGB stars with different masses, namely: the third dredge-up – that determines the increase of the surface carbon abundance –, and mass loss by stellar winds, that affects the duration, hence the luminosity excursion during this phase. Iben (1981) first pointed out the so-called "carbon star mystery", corresponding to a long-standing discrepancy between theory and observations, i.e. current stellar mod- els predicted a deficit of faint carbon stars, accompanied by an excess of bright carbon (in general AGB) stars. This situation is exemplified in Fig. 10, where we report the pre- dicted CSLF for the LMC, according to the calculations performed by Iben & Renzini (1983), with model prescrip- tions very similar to RV81. For this particular case, the authors adopt the following set of parameters: efficiency parameter η = 2/3 in the Reimers (1975) mass-loss for- mula; mixing-length parameter, α = 2; minimum core for the third dredge-up to occur, M min = 0.5M⊙; and effi- ciency of the third dredge-up, λ, as a function of the core mass. c The CSLF in the LMC is instead very well fitted (Fig. 10) by the other two AGB synthetic models here con- sidered, namely: Groenewegen & de Jong (1993, GdJ93; top right-hand side panel) and Marigo et al. (1999a, MGB99; bottom left-hand side panel). We remind again that, differently from RV81, in these studies the third dredge-up is suitably calibrated in order to reproduce the CLSF in the LMC. Finally, it should be remarked that Marigo et al. (1999a) have extended the analysis to the CSLF in the SMC (see bottom right-hand side panel of Fig. 5.2.3), so as to include a metallicity-dependent treat- ment of the third dredge-up in their synthetic AGB model. 5.2.4. The chemical abundances of planetary nebulae The observed chemical composition of planetary nebulae represents another crucial constraint for stellar models, as it is the record of the nucleosynthetic and mixing processes occurred during the previous stellar evolution. As far as HG97 predictions are concerned, a detailed discussion is given in Groenewegen & de Jong (1994) and it will not be repeated here. We restrict here to the results of RV81 and M2K which are compared in Fig. 11 with the measured abundances of He, C, and O in galactic plane- tary nebulae. Predicted PN abundances can be found in Tables A13 – A15. Since a full analysis is beyond the pur- pose of this work, we simply consider the most relevant aspects. Both RV81 and M2K results shown here are derived from calculations of stellar models with initial solar com- position. Therefore, they cannot reproduce the data points with He/H <∼ 0.10 − 0.11, since these latter most likely correspond to progenitor stars with initial subsolar metal- licity. The expected paths of PN abundances as a function of the initial mass of the progenitor star reflect the efficiency and duration of the involved processes. For instance, in both models the C/O ratio first increases at increasing mass due to the third dredge-up, and then (for M > 3 − 4M⊙) it starts to decline because of HBB. An interesting point is the anticorrelation between the N/O ratio and the C/O ratio exhibited by observed PNe with the highest helium content (He/H> 0.125; the so- called type I PNe according to the classification intro- duced by Peimbert (1978)). This trend is reproduced by M2K synthetic calculations, tracing the signature of HBB in the most massive AGB stars (M >∼ 3.5M⊙), where car- bon in the envelope is quickly converted into nitrogen. Note that theoretical results are notably sensitive to the adopted value for the mixing-length parameter, i.e. the larger α is, the more efficiently HBB operates, yield- ing higher N/O and lower C/O ratios. Limiting to the observed sample of PNe, we might deduce that HBB has operated in the most massive progenitors of solar metal- licity, but with a rather mild efficiency, in agreement with the conclusion already mentioned by Henry et al. (2000a). In fact, as we can see from Fig. 11, predictions for the case (α = 2.5) – which correspond to strong HBB – lead to N/O ratios quite higher than observed. Hower, these indications should be considered with some caution, as the considered sample of PNe might not cover the whole relevant mass range of the progenitors (see Henry et al. 2000a), and predictions of PNe abundances are derived under very simple assumptions (see Appendix A). A much better approach will be adopted with the aid of a detailed synthetic model of PN evolution, which is being developed (Marigo et al. 2001, in preparation; see Marigo et al. 1999b for a preliminary presentation). Finally, we would remark that RV81 results for the most massive stellar models – shown in Fig. 11 with dotted lines – do not correspond to PN abundances, but rather to surface abundances just prior the progenitor stars ex- plode as SNe I-1/2. Therefore, attention must be paid not to use these data for a comparison with observed PN abun- dances. 5.3. Yields from simple stellar generations We will compare here the stellar yields presented in this work with those calculated by RV81 and HG97. Possi- ble differences are then discussed on the basis of different 16 P. Marigo Fig. 11. Abundance ratios of galactic planetary nebulae. Observed data (squares) are taken from Kingsburgh & Barlow (1994), and Henry et al. (2000a). Filled squares correspond to PNe with log(N/O) > −0.5 and He/H > 0.125. Predicted PN abundances are shown as a function of the initial mass of solar-metallicity progenitor stars, as derived in this work (left-hand side panel), and in Renzini & Voli (1981, right-hand side panel). Lines (solid and dashed) connect predicted abundances at increasing stellar mass (a few values are indicated nearby) for two values of the mixing length parameter. In the case of RV81 model, surface abundances just prior Chandrasekhar carbon explosion are also shown (dotted line). model prescriptions and observational constraints, already mentioned in the previous sections. We choose not to make a direct comparison between yields produced by models with the same initial mass, be- cause different sets of yields cover different mass-ranges in the domain of low- and intermediate-mass stars (see Sect. 5.1.1). For this reason, it is more meaningful to com- pare the whole chemical contribution provided by low- and intermediate-mass stars belonging to a given simple (i.e. coeval) stellar population. To this aim, we recall that ac- cording to the standard definition (Tinsley 1980), the yield from a stellar generation, yk, is the mass converted into the chemical element k and ejected by all stars per unit mass locked into stars: where φ(M ) = dN/dMi is the IMF (by number) de- fined between the lower (Ml) and upper (Mu) mass limits; W (Mi) is the remnant mass. In order to weigh the sole contribution from the gener- ation of low- and intermediate-mass stars, let us consider the quantity: ylims Ml k = R Mup R Mu Ml M pk(Mi)φ(Mi)dMi Miφ(Mi)dMi (13) that is similar to Eq. (11) with Mu = Mup. The adopted integration extremes are Ml = 0.9M⊙, Mu = 100M⊙, and Mup according to the set of stellar yields under consider- ation (see Table 1 and Sect. 5.1.1). The IMF is expressed by the classical Salpeter's law, i.e. φ(Mi) = dN/dMi ∝ M −(1+x) with x = 1.35 (Salpeter 1955). k The quantities ylims express the relative chemical con- tribution (for a given elemental species k) from low- and intermediate-mass stars belonging to a given simple stel- lar population. They are shown in Fig. 12 as a function of the metallicity for the three sets here considered. It should be remarked that the differences between our results (M2K) and those derived by RV81 and HG97 are not only due to the different mass-range covered by low- and intermediate- mass stars, but mainly reflect substan- yk = (1 − R)R Mu Ml Mipk(Mi)φ(Mi)dMi Miφ(Mi)dMi R Mu Ml (11) i In the above equation pk(m) is the stellar yield of the kth element (see Sect. 3), and R is the returned fraction, expressing the fraction of mass that has formed stars and then been ejected: R = R Mu Ml [Mi − W (Mi)]φ(Mi)dMi Miφ(Mi)dMi R Mu Ml (12) Yields from low- and intermediate-mass stars 17 Fig. 12. Integrated yield contribu- tions from low- and intermediate- mass stars as a function of the metallicity, as defined by Eq.(12). The mixing-length parameters (α) adopted by the authors are indi- cated. tial differences in the adopted physical prescriptions as already illustrated in Sect. 5.1. Differences essentially show up both in metallicity trends and absolute values of ylims . Compared to previ- ous calculations, M2K yields show a pronounced depen- dence on the metallicity, i.e. positive yields increase with decreasing Z. Conversely, the RV81 and HG97 sets present weak trends with Z. k The metallicity dependence can be explained as fol- lows. On one side, AGB lifetimes of low-mass stars in- crease at decreasing metallicities, as mass-loss rates are expected to be lower. This fact leads to a larger num- ber of dredge-up episodes. Moreover, both the onset and the efficiency of the third dredge-up are favoured at lower metallicities. These factors concur to produce a greater enrichment in carbon. On the other side, HBB in more massive AGB stars becomes more efficient at lower metal- licities, leading to a greater enrichment in nitrogen. The combination of all factors favours higher positive yields of helium at lower Z. As far as the single elemental species are concerned, we can notice: – M2K yields of 4He are larger than those by HG97 and RV81 towards lower Z, likely due to the earlier activa- tion and larger efficiency of the third dredge-up in our models. With respect to RV81 and HG97 predictions, our yields of 4He present a significant trend with Z. – M2K yields of 12C are systematically higher than those of RV81 and HG97 because of the earlier onset (and average greater efficiency than in RV81) of the third dredge-up. – The dominant contribution to the yields of 14N comes from HBB in the most massive AGB stars. Differences in the results reflect different efficiencies of nuclear re- actions and AGB lifetimes. In particular, according to M2K the production of 14N, mainly of primary synthe- sis, is favoured at lower Z, and is sistematically lower than RV81 and HG97 results. This latter difference can be explained considering the drastic reduction of TP-AGB lifetimes for the most massive AGB models with HHB (see Fig. 6) in M2K models with respect to RV81. In general, our expected dependence of chemical yields on metallicity is far for being linear, and much caution should be used when extrapolating these quantities with respect to Z in chemical evolutionary models. We cannot verify whether such a non-linear relation with metallicity was displayed also in the RV81 models, since just two val- ues of metallicities were considered there. However, the very high number of thermal pulses suffered by stars with 18 P. Marigo HBB regardless of the metallicity, according to the RV81 models, might already explain the apparent lower sensi- tiveness of their yields to the metallicity. 6. Final remarks We would like to outline briefly the aims and findings of the present work. In the first part we have presented new homogeneous sets of stellar yields ejected from low- and intermediate- mass stars, in view of providing updated ingredients for modelling the chemical evolution of complex stellar sys- tems. Thanks to the updated input physics employed in the calculations, and the improved treatment of both the third dredge-up and hot-bottom burning, the present es- timation of the stellar yields from low- and intermediate- mass stars has led to new results and developments. In particular, a pronounced trend of the yields with the metallicity is expected. Specifically, at given stellar mass positive yields of 4He, 12C, 14N, and 16O are larger at decreasing metallicity. This feature is the result of con- curring factors: at lower Z both the third dredge-up and hot-bottom burning are more efficient, and TP-AGB life- times are, on average, longer because of lower mass-loss rates. Moreover, it is interesting to notice that low-mass stars may produce positive yields of 16O, which is brought up to the surface by the third dredge-up. The entity of this con- tribution (as well as for that of 12C) depends crucially on the efficiency λ, number of dredge-up episodes, and chemi- cal composition of the convective intershell (for this latter, see Boothroyd & Sackmann 1988b, and Herwig 2000 for different model results). The possible production and ejection of newly synthe- sised oxygen from low- and intermediate-mass stars may be an interesting prediction to be tested through its con- sequences in various possible applications, mostly in rela- tion to the chemical composition of planetary nebulae (e.g. P´equignot et al. 2000), and galactic chemical evolutionary models. The second part of the paper is meant to examine sev- eral aspects concerning the stellar models which the chem- ical yields are derived from. To this aim, we have analysed how the main model prescriptions (i.e. treatment of con- vective boundaries, mass loss, analytical relations, third dredge-up parameters, treatment of hot-bottom burning, etc.) may affect the predictions of stellar yields. Finally, the third final part is dedicated to compare our new yields with other available data sets of large usage, in the attempt of explaining the existing differences as the result of particular assumptions. To do this, we have con- sidered basic observational constraints which are closely related to the stellar yields – namely: i) the initial-final mass relation; ii) the white dwarf mass distribution; iii) the carbon star luminosity function; and iv) the chemical composition of planetary nebulae – and tested the capa- bility of different models in reproducing them through a direct comparison between predictions and observations. In particular, it has been shown how much the choice of calibrating fundamental efficiency parameters (e.g. of the third dredge-up and mass loss) in recent works has changed the predictions of stellar yields compared to ear- lier studies. To conclude, we wish this work has somehow con- tributed to clarify a few important general points on the- oretical stellar yields, in view of stimulating an aware and critical usage of them. Acknowledgements. I would like to thank L´eo Girardi, Laura Portinari, and Cesare Chiosi for their important professional advice and support, and the referee, Dr. R.B.C. Henry, for his helpful remarks on this work. This study has been financially supported by the Italian Ministry of University, Scientific Re- search and Technology (MURST) under contract "Formation and evolution of Galaxies" n. 9802192401. References Alongi M., Bertelli G., Bressan A., Chiosi C., Fagotto F., Greg- gio L., Nasi E., 1993, A&AS 97, 851 Baud B., Habing H.J., 1983, A&A 127, 73 Bergeron P., Saffer R.A., Liebert J., 1992, ApJ 394, 228 Blocker T., 1995, A&A 297, 727 (B95) Blocker T., Schonberner D., 1991, A&A 244, L43 Boothroyd A.I., Sackmann I.-J., 1988a, ApJ 328, 641 Boothroyd A.I., Sackmann I.-J., 1988b, ApJ 328, 653 Boothroyd A.I., Sackmann I.-J., 1999, ApJ 510, 232 Bowen G.H., 1988, ApJ 329, 844 Bragaglia A., Renzini A., Bergeron P., 1995, ApJ 443, 735 Caughlan G.R., Fowler W.A., 1988, Atomic Data Nucl. Data Tables 40, 283 Charbonnel C., 1995, ApJ 453, L41-L44 Chiosi C., Bertelli G., Bressan A., 1992, ARA&A 30, 305 Costa E., Frogel J.A., 1996, AJ 112, 2607 Crotts, A.P.S., Bechtold J., Fang Y., Duncan R.C., 1994, Bull. American Astron. Soc. 185, 1807 Dominguez I., Chieffi A., Limongi M., et al., 1999, ApJ 524, 226 Fleischer A.J., Gauger A., Sedlmayr E., 1992, A&A 266, 321 Forestini M., Charbonnel C., 1997, A&AS 123, 241 Garcia-Berro E., Ritossa C., Iben I., 1997, ApJ 485, 765 Girardi L., Bressan A., Bertelli G., Chiosi C., 2000, A&AS 141, 371 Graboske H.C., de Witt H.E., Grossman A.S., Cooper M.S., 1973, AJ 181, 457 Gratton R.G., Sneden C., Carretta E., Bragaglia A., 2000, A&A 354, 169 Groenewegen M.A.T., de Jong T., 1993, A&A 267, 410 Groenewegen M.A.T., de Jong T., 1994, A&A 282, 115 Henry R.B.C., Kwitter K.B., Bates J.A., 2000a, ApJ, 531, 928 Henry R.B.C., Edmunds M.G., Koppen J., 2000b, ApJ, 541, 660 Herwig F., 1996, in Stellar Evolution: What Should Be Done, 32nd Li`ege Int. Astrophys. Coll., eds. A. Noels et al., p. 441 Herwig F., Blocker T., Schonberner D., El Eid M., 1997, A&A 324, L81 Yields from low- and intermediate-mass stars 19 Appendix A: Tables of stellar yields and planetary nebulae chemical abundances Tables A1 – A12 contain the yields from low- and intermediate-mass stars, in the form My (see Eq. (1)), as a function of the initial mass Mi and metallicity Z. All specified masses are expressed in solar units. Concerning low-mass stars (with Mi <∼ MHeF), we give separately the yields ejected during the RGB (Ta- bles A1 – A3) and AGB phases (Tables A4 – A6). The quantity, MTP,0, corresponds to the mass at the onset of the TP-AGB phase, which is smaller than Mi by the amount of mass lost during the previous RGB phase, ∆Mej(RGB). The mass lost during the AGB is denoted with ∆Mej(AGB). Total yields produced by stars in the whole mass range (0.8M⊙ <∼ Mi <∼ 5M⊙) are presented in Tables A7 – A12. The total amount of ejected mass is denoted with ∆Mej. Total yields from stars with Mi ≥ 3.5M⊙ are given in Tables A10 – A12 for three values of the mixing-length parameter α, and distinguishing between the secondary (S entry) and primary (P entry) components in the case of the CNO elements. Tables A13 – A15 present the predicted abundances ratios (He/H, C/H, N/H, O/H) in planetary nebulae, as a function of the initial stellar mass (Mi), metallicity Z, and mixing-length parameter. The PNe chemical abundances (by number, in mole gr−1) are calculated by averaging the abundances in the wind ejecta over the last stages (i.e. a time period τPN = 3 × 104 yr) on the AGB, weighted by the masses of the ejecta. For the sake of simplicity, we do not consider the ques- tion on the actual observability of PNe (which depends on both dynamical and ionisation properties), and the fact that evolutionary timescales of PNe (hence the time τPN) may largely vary according to the mass of the progenitor star. These points deserve a more complex study, which is currently in progress and presented in a preliminary form by Marigo et al. (1999b). All data are available in electronic format at the web site address: http://pleiadi.pd.astro.it. Herwig F., 2000, A&A 360, 952 van den Hoek L.B., Groenewegen M.A.T., 1997, A&AS 123, 305 (HG97) Iben I., 1981, ApJ 246, 278 Iben I., Renzini A., 1983, ARA&A 21, 27 Iben I., Truran J.W., 1978, ApJ 220, 980 (IT78) Jeffries R.D., 1997, MNRAS 288, 585 Jeffries R.D., James D.J., Thurston M.R., 1998, MNRAS 300, 550 Kingsburgh R.L., Barlow M.J., 1994, MNRAS 271, 257 Koester D., Reimers D., 1996, A&A 313, 810 Lu, L.. 1991, ApJ 379, 99 Lu, L., Sargent W.L.W., Barlow T.A., 1998, AJ 115, 55 Maeder A., 1983, A&A 120, 113 Maeder A., Meynet G., 2000, ARA&A 38, in press Marigo P., 1998a, PhD Thesis, University of Padova Marigo P., 1998b, A&A 340, 463 Marigo P., Bressan A., Chiosi C., 1996, A&A 313, 545 Marigo P., Bressan A., Chiosi C., 1998, A&A 331, 564 Marigo P., Girardi L., Bressan A., 1999a, A&A 344, 123 Marigo P., Weiss A., Groenewegen M.A.T., Girardi L., 1999b, in From extrasolar planets to Cosmology: The VLT Open- ing Symposium, Proceedings ESO/MPA Parallel Work- shop 2: Star-Way to the Universe, held in Antofagasta (Chile), March 1999, Eds. Bergeron J., Renzini A., p. 248- 251 Peimbert M., 1978, in IAU Symposium 76, Planetary Nebulae, ed. Y. Terzian, Dordrecht: Reidel, p. 215 P`equignot D., Walsh J.R., Zijlstra A.A., Dudziak G., 2000, A&A 361, L1 Pettini M., Lipman K., Hunstead R.W., 1995, ApJ 451, 100 Portinari L., Chiosi C., Bressan A., 1998, A&A 334, 505 Prantzos, N., Vangioni-Flam E., Casse M., 1994, RAS Canada J. 88, 356 Rebeirot E., Azzopardi M., Westerlund B.E., 1993, A&AS 97, 603 Reid, N., 1996, AJ 111 (5), 2000 Reimers D., 1975, Mem. Soc. R. Sci. Li`ege, ser. 6, vol. 8, p. 369 Renzini A., Voli M., 1981, A&A 94, 175 (RV81) Salpeter E.E., 1955, ApJ 121, 161 Tinsley B.M., 1980, Fundam. of Cosmic Phys. 5, 287 Tytler D., Fan X.-M., 1994, ApJ 424, L87 Vassiliadis E., Wood P.R, 1993, ApJ 413, 641 Vila-Costas M.B., Edmunds M.G., 1993, MNRAS 265, 199 Wagenhuber J., Groenewegen M.A.T., 1998, A&A 340, 183 Wasserburg G.J, Boothroyd A.I., Sackmann I.-J., 1995, ApJ 447, L37 Weidemann V., 1987, A&A 188, 74 Weidemann V., 1997, in Advances in Stellar Evolution, Pro- ceedings of the Workshop Stellar Ecology, Cambridge Uni- versity Press, p. 169 Weiss A., Denissenkov P.A., Charbonnel A., 2000, A&A 355, 299 Woosley S.E., Weaver T.A., 1982, in Supernovae: a Survey of Current Research, Reidel, Dordrecht, eds. Rees M. J., Stoneham R. J., p.79 Wood, P.R., 1981, in: Physical processes in red giants, Proc. of the Second Workshop, Erice, Italy (September 3-13, 1980), Dordrecht, D. Reidel Publishing Co., p. 135-139 20 Appendix A: TABLES OF STELLAR YIELDS Table A. STELLAR YIELDS EJECTED DURING THE RGB PHASE { INITIAL METALLICITY Z = : M (cid:1)M (RGB) M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i ej y y y y y y y y y y y          . .E- -. E- .E- . E- -.E- . E- . E- -.E- -. E- .E- -.E- . E- . .E- -. E- .E- . E- -.E- . E- .E- -.E- -.E- .E- -.E- . E- .  .E- -.E- . E- .E- -.E- .E- .E- -. E- -.E- . E- -.E- .E- . .E- -.E- .E- . E- -.E- . E- .E- -.E- -.E- .E- -.E- . E- . .E- -.E- .E- . E- -.E- .E- . E- -. E- -.E- . E- -.E- .E- . .E- -.E- .E- .E- -.  E- . E- . E- -.E- -.E- . E- -.E- .E- . .E- -.E- .E- .E- -. E- .E- .E- -.E- -.E- . E- -.E- .E- . .E- -.E- .E- .E- -.  E- .E- .E- -. E- -.E- .E- -. E- .E- . .E- -.E- .E- . E- -.E- .E- .E- -. E- -. E- . E- -.E- .E- . .E- -.E- .E- .E- -.E- .E- . E- -. E- -.E- .E- -. E- . E- . .E- -.E- .E- .E- -. E- .E- .E- -.E- -.E- .E- -.  E- .E- . .E- - .E- .E- .E- -.E- . E- .E- -.E- -.E- .E- -. E- . E- . .E- -.E- . E- .E- -.E- . E- . E- - .E- -.E- .E- -.E- .E- . . E- -.E- .E- .E- -.E- .E- .E- -.E- -.E- .E- -.E- .E- .  .E- -.E- .E- .E- -.E- .E- .E- -.E- -. E- .E- - .E- .E- Table A. STELLAR YIELDS EJECTED DURING THE RGB PHASE { INITIAL METALLICITY Z = : M (cid:1)M (RGB) M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i ej y y y y y y y y y y y          . .E- -.E- .E- . E- -.E- .E- .E- -. E- -.E- .E- -.E- .E- . .E- -.E- .E- .E- -.E- .E- .E- -.E- -.E- .E- -.E- . E- .  .E- -. E- .E- .E- -. E- .E- . E- -.E- - . E- .E- -.E- .E- .  . E- -.E- . E- .E- -.E- .E- .E- -. E- -.E- .E- -. E- .E- . .E- -.E- . E- .E- -.E- .E- .E- -.E- -.E- .E- -.E- .E- . .E- -.E- .E- .E- -.E- .E- . E- -.E- -.E- . E- -.E- .E- . . E- -. E- .E- . E- -. E- .E- . E- -.E- -.E- .E- - . E- .E- . .E- -.E- .E- .E- -.E- .E- .E- -.E- -.E- .E- - .E- .E- . .E- -.E- .E- .E- -.E- .E- .E- -.E- -. E- .E- -.E- .E- . . E- -.E- . E- .E- -.E- . E- . E- -.E- -.E- .E- -. E- .E- . .E- -.E- . E- .E- -.E- .E- .E- -. E- -.E- .E- -.E- . E- . .E- -.E- .E- .E- -. E- . E- .E- -. E- -. E- . E- -.E- .E- . .E- -.E- .E- .E- -. E- .E- . E- -. E- -.E- .E- -.E- . E- . .E- -.E- .E- .E- -. E- .E- .E- -.E- -.E- .E- -. E- .E- Table A. STELLAR YIELDS EJECTED DURING THE RGB PHASE { INITIAL METALLICITY Z = : M (cid:1)M (RGB) M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i ej y y y y y y y y y y y          . .E- -.E- .E- . E- -.E- .E- .E- -.E- -.E- . E- -.E- .E- . .E- -.  E- . E- .E- -. E- .  E- .E- -.E- -.E- .E- -. E- . E- .  .E- -.E- .E- .E- -.E- . E- . E- -. E- -.E- .E- -.E- .E- . .E- -.E- .E- .E- -.E- . E- .E- -. E- -.E- .E- -.E- .E- . .E- -.E- .E- .E- -. E- . E- .E- -.E- -.E- .E- -.E- . E- . .E- -.E- .E- .E- -.E- .E- . E- -. E- -.E- .E- -. E- .E- . .E- -.E- .E- .E- -.E- .E- . E- -. E- -. E- .E- -. E- .E- .  .E- -.E- .E- .E- -.E- . E- .E- -.E- -.E- .E- -.E- .E- . .E- -.E- .E- . E- -.E- . E- . E- -.E- -.E- .E- -.E- .E- . .E- - . E- .E- .E- -.E- .E- . E- -. E- -. E- .E- -.E- .E- . .E- -. E- .E- . E- - . E- .E- .E- -.E- -.E- .E- -.E- .E- . . E- -. E- .E- . E- -. E- .E- .E- -.E- -.E- .E- -. E- .E- Table A. STELLAR YIELDS EJECTED DURING THE AGB PHASE { INITIAL METALLICITY Z = : M M (cid:1)M (AGB) M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i TP; ej y y y y y y y y y y y          . . .E- -.E- .E- .E- -.E- .  E- .E- -.E- -.E- .E+ -. E- .E- . . .E- -. E- .E- .E- -.E- .E- . E- -.E- -.E- .E- -.E- . E- .  . .E- -.E- .E- .E- -.E- .E- .E- -.E- -.E- . E- -.E- .E- . . .E- -.E- . E- . E- -. E- .E- . E- -.E- -.E- .E- -.E- .E- . . .E- -.E- .E- .E- -. E- .E- .E- -.E- -.E- . E- -. E- .E- . . .E- -.E- .E- . E- -.E- . E- .E- -. E- -.E- .E- -.E- .E- . . . E- - .E- .E- .E- -.E- .E- .E- -.E- -.E- .E- -. E- .E- . . .E- -. E- .E- .E- -.E- . E- .E- -.E- -.E- .  E- -.E- .E- . . .E- -. E- .E- . E- -.E- . E- .E- -.E- -.E- .E- -.E- .E- . . .E- -. E- .E- . E- .E- .E- .E- -. E- .E- .E- . E- .E- . . . E- -. E- .E- .E- .E- .E- .E- -.E- . E- .E- -.E- .E- . . .E+ -.E- . E- . E- .E- .E- .E- -. E- .E- .E- -.E- .E- . . .E+ -.E- .E- .E- .E- .E- .E- -. E- .E- . E- -.E- .E- . . .E+ -.E- .E- . E- .E- . E- .E- -.E- .E- .E- -.E- .E- .  . .E+ -. E- . E- .E- .E- .E- .E- -.E- .E- .E- -.E- .E- Appendix A: TABLES OF STELLAR YIELDS 21 Table A. STELLAR YIELDS EJECTED DURING THE AGB PHASE { INITIAL METALLICITY Z = : M M (cid:1)M (AGB) M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i TP; ej y y y y y y y y y y y          . . .E- -.E- .E- . E- -.  E- .E- .E- -. E- -.E- .E+ -.E- .E- .  . .E- -. E- .E- .E- -.E- .E- .E- -. E- -.E- .E- - .E- .E- .  . .E- -.E- . E- .E- -.E- .E- .E- -. E- -.E- .E- -. E- .E- . . .E- -.E- .E- . E- -.E- .E- . E- - . E- -.E- . E- -. E- .E- . . .E- -.E- .E- . E- -.E- .E- . E- -.E- -.E- .E- -.E- .E- . . .E- -.E- . E- .E- -.E- . E- .E- -.E- -.E- . E- -.E- .E- . . . E- -.E- .E- .E- .E- .E- .E- -.E- .E- .E- -.E- .E- . . . E- -. E- .E- .E- .E- .E- .E- -. E- .E- . E- -.E- . E- . . . E- -.E- . E- .E- .E- .E- .E- -. E- .E- .E- -. E- .E- . . .E- -.E- .E- .E- .E- .E- .E- -.E- .E- .E- -.E- .E- . . .E- -.E- .E- .E- .E- .E- .E- -.E- .E- .E- -.E- . E- . . .E+ -.E- .E- .E- . E- .E- .E- -.E- .E- . E- -. E- .E- . . .E+ - .E- . E- . E- .E- .E- . E- - .E- .E- .E- -.E- .E- Table A. STELLAR YIELDS EJECTED DURING THE AGB PHASE { INITIAL METALLICITY Z = : M M (cid:1)M (AGB) M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i TP; ej y y y y y y y y y y y          . . .E- -. E- .E- . E- - .E- . E- .E- -.E- -.E- .E- -. E- .E- . . .E- -. E- .E- .E- -.E- .E- .E- -.E- -.E- .E- -.E- .E- .  . . E- -.E- .E- . E- -.E- .E- . E- -.E- -.E- .E- -.E- .E- . . .E- -.E- .E- . E- -.E- .E- . E- -. E- -.E- .E- -. E- . E- . . .E- -. E- .E- .E- -.E- . E- .E- -.E- -.E- .E- -.E- . E- . . .E- -.E- .E- . E- . E- .E- .E- -.E- .E- .E- -. E- .E- . . .E- -.E- .E- .E- .E- . E- .E- -. E- .E- .E- -.E- .E- .  . . E- - . E- .E- .E- .E- .E- .E- -.E- .E- .E- -. E- .E- . . .E- -.E- . E- . E- . E- .E- .E- -.E- .E- .E- - .E- .E- . . . E- -.E- .E- . E- . E- .E- .E- -.E- .E- .E- -.E- .E- . . .E- -.E- . E- .E- . E- .E- .E- -.E- . E- . E- -.E- .E- . . .E+ -. E- .E- .E- .E- .E- .E- -. E- .E- .E- -. E- .E- Table A. TOTAL STELLAR YIELDS { MIXING-LENGTH PARAMETER (cid:11) = : { INITIAL METALLICITY Z = : M (cid:1)M M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i ej y y y y y y y y y y y          . . E- -. E- . E- .E- -.E- . E- . E- -.E- -.E- .E- -. E- .E- . .E- -.E- .E- .E- -.E- .E- .E- -. E- -. E- .E- -.E- . E- .  . E- -.E- .E- .E- -.E- .E- .E- -.E- -. E- . E- -. E- .  E- . .E- -.E- . E- . E- -. E- .E- .E- -.E- -.E- .E- -.E- .E- . .E- -. E- . E- .E- -.E- .E- . E- -. E- -. E- .E- -.E- .E- . .E- - . E- .E- .E- -.E- .E- .E- -. E- -.E- .E- -.E- . E- . .E- -.E- .E- . E- -.E- .E- .E- -.E- -.E- .E- -.E- .E- . .E- - .E- .E- . E- -.E- .E- .E- -.E- -.E- .E- -.E- . E- . .E- -.E- .E- .E- -.E- .E- .E- - .E- -.E- .E- -.E- .E- . .E- -.E- . E- .E- .E- . E- .E- -. E- .E- . E- -.E- . E- . .E+ -.E- .E- .E- .E- . E- .E- -.E- .E- .E- -.E- .E- . .E+ -.E- .E- .E- .E- .E- .E- -. E- .E- .E- -.E- .E- . .E+ -.E- . E- .E- .E- .E- . E- -.E- .E- .E- -.E- .E- . .E+ -. E- .E- .E- .E- . E- .E- -.E- . E- . E- -.E- .E- .  .E+ -.E- .E- .E- .E- .E- .E- -.E- .E- .E- -.E- .E- . .E+ -.E- . E- .E- .E- .E- .E- -.E- .E- .E- -. E- .E- . . E+ -.E- . E- .E- .E- .E- . E- -.E- . E- .E- -.E- .E- . .E+ -.E- . E- .E- .E- .E- .E- -.E- .E- .E- -.E- .E- . .E+ -.E- .E- .E- .E- .E- .E- -.E- .E- .E- -.E- .E- Table A. TOTAL STELLAR YIELDS { MIXING-LENGTH PARAMETER (cid:11) = : { INITIAL METALLICITY Z = : M (cid:1)M M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i ej y y y y y y y y y y y          . .E- -.E- .E- .E- -.E- . E- .E- -.E- -.E- .E- -.E- . E- .  . E- -.E- .E- . E- -. E- . E- .E- -. E- -.E- .E- -.E- . E- .  .E- -. E- . E- .E- -.E- .E- .E- -.E- -.E- .E- -.E- .E- . . E- -. E- .E- .E- -.E- .E- .E- -.E- -.E- .E- -.E- .E- . .E- - . E- .E- .E- -. E- . E- .E- -. E- -. E- .E- -. E- . E- . .E- -.E- .E- . E- -.E- .E- .E- -. E- -.E- . E- -.E- .E- . .E- -.E- .E- . E- .E- . E- .E- -.  E- .E- .E- -.E- .E- . . E- -. E- .E- .E- . E- .E- .E- -.E- .E- .E- -.E- . E- . . E- -. E- . E- .E- .E- .E- .E- -.E- .E- . E- -.E- .E- . .E- -.E- .E- . E- .E- .E- .E- -. E- .E- . E- -. E- .E- . .E+ -.E- .E- .E- .E- .E- .E- -.E- .E- .E- -.E- . E- . .E+ -.E- .E- .E- . E- . E- .E- - .E- .E- .E- -. E- .E- . .E+ - . E- .E- .E- .E- . E- . E- - . E- .E- .E- -.E- .E- . . E+ -.E- . E- .E- .E- . E- .E- -.E- .E- . E- -. E- .E- . .E+ -.E- .E- .E- . E- . E- . E- -.E- .E- .E- -.E- .  E- . .E+ -. E- .E- . E- .E- .E- .E- -.E- . E- .E- -.E- .E- . . E+ -.E- .E- .E- .E- .E- . E- -.E- .E- .E- -. E- .E- . .E+ -. E- .E- .E- . E- .E- .E- -.E- .E- .E- -.E- .E- . .E+ -.E- . E- .E- . E- .E- .E- -.  E- . E- .E- -. E- .E- 22 Appendix A: TABLES OF STELLAR YIELDS Table A . TOTAL STELLAR YIELDS { MIXING-LENGTH PARAMETER (cid:11) = : { INITIAL METALLICITY Z = : M (cid:1)M M (H) M ( He) M ( He) M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i ej y y y y y y y y y y y          . .E- -.E- .E- .E- -. E- . E- .E- -. E- -. E- .E- -.E- .E- . .E- -.E- . E- .E- - . E- .E- . E- -.E- -.E- .E- -.E- .E- .  .E- -.E- .E- .E- -. E- . E- .E- -.E- -.E- . E- -. E- . E- . .E- -. E- . E- .E- -.E- .E- .E- -.E- -.E- .E- -. E- .E- . . E- -.E- .E- . E- -.E- .E- .E- -.E- -.E- .E- -.E- .E- . .E- -.E- . E- . E- .  E- .E- .E- -.E- .E- .E- -.E- .E- . .E- -. E- . E- .E- .E- .E- . E- -.E- .E- .E- -. E- .E- .  .E- - .E- .E- .E- .E- .E- .E- -.E- .E- . E- -. E- .E- . .E- -.E- .E- .E- . E- .E- .E- -.E- .E- .E- - .E- .E- . .E- -. E- .E- . E- . E- . E- .E- -.E- .E- .E- -.E- .E- . . E- -.E- . E- .E- . E- .E- .E- -.E- . E- . E- -.E- .E- . .E+ -.  E- .E- .E- .E- . E- .E- -.E- .E- .E- -.E- .E- . . E+ -.E- . E- .E- . E- .E- .E- -. E- .E- .E- -.  E- .E- . .E+ -.E- . E- .E- .E- .E- .E- -. E- .E- .E- -.E- . E- . .E+ -.E- .E- .E- .E- .E- .E- -.E- .E- .E- -.E- . E- . .E+ -. E- .E- .E- . E- . E- .E- -.E- .E- .E- -.E- .E- . . E+ -.E- .E- . E- .E- .E- .E- -.E- .E- .E- -.E- .E- . .E+ -.E- . E- .E- .E- .E- .E- -.E- .E- .E- -.E- . E- Table A. TOTAL STELLAR YIELDS { INITIAL METALLICITY Z = : (cid:11) M (cid:1)M M (H) M ( He) M ( He) T/S/P M ( C) M ( C) M ( N) M ( N) M ( O) M ( O) M ( O) M (CNO) i ej y y y y y y y y y y y          . . .E+ -.E- .E- .E- T .E- .E- .E- -.E- -.E- .E- -.E- .E- S -.E- .E- .E- -.E- -.E- .E- -.E- -.E- P . E- . E- .E- -.E- . E- . E- .E- .E- . . E+ -.E- . E- .E- T .E- .E- .E- -.E- -.E- .E- -.E- .E- S -.E- .E- .E- -.E- -. E- .E- -.E- - .E- P . E- .E- .E- -.E- .E- .E- .E- .E- . .E+ -. E- .E- .E- T .E- .E- . E- -.E- -.E- . E- -. E- . E- S -.E- .E- . E- -.E- -.E- .E- -. E- -.E- P . E- .E- -.E- -.E- .E- .E- -.E- .E- . .E+ -.E- -.E- .E- T .E- .E- .E- -.E- -. E- . E- -.E- .E- - S -.E- .E- .E- -.E- -. E- . E- -.E- -.E- P .E- .E- . E- .E- .E- .E- . E- .E- . . . E+ -. E- . E- .E- T . E- .E- . E- -.E- . E- . E- -.E- .E- S -. E- .E- . E- -.E- -.E- . E- -.E- -.E- P .E- . E- .E- .E- .E- -.E- -.E- . E- . .E+ -. E- .E- .E- T .E- . E- . E- -.E- -.E- .E- -.E- . E- S -.E- .E- .E- -.E- -.E- .E- -.E- -. E- P .E- .E- .E- .E- .E- .E- -.E- .E- . .E+ -.E- -.E- . E- T -.E- .E- . E- -.E- -.E- .E- -.E- .E- . . E+ -. E- -. E- . E- T - .E- .E- .E- -.E- -. E- .E- -. E- .E- S -.E- .E- .E- -.E- -.E- .E- -.E- -.E- P .E- . E- . E- .E- . E- .E- .E- .E- S -. E- .E- . E- -. E- -.E- . E- -. E- .E- P . E- .E- .E- .E- .E- . E- . E- .E- . . .E+ -. E- -.E- .E- T .E- . E- .E- - . E- .E- . E- -.E- . E- . .E+ -.E- -.E- .E- T -.E- . E- . E- -.E- -.E- .E- -.E- .E- S -.E- .E- .E- - .E- -. E- . E- -.E- -.E- P .E- .E- . E- .E- .E- .E- . E- .E- . .E+ -. E- -.E- .E- T -.E- . E- .E- -.E- -. E- . E- -.E- . E- . . E+ -.E- -.E- .E- T -.E- .E- . E- -. E- -.E- .E- -. E- . E- S -.E- .E- .E- -.E- -.E- .E- -.E- -.E- P . E- .E- .E- . E- .E- .E- . E- . E- S - .E- .E- . E- -.E- -.E- .E- -. E- -. E- P .E- .E- . E- .E- .E- .E- . E- .E- S -.E- . E- .E- -.E- - .E- . E- -. E- . E- P . E- .E- . E- .E- .E- .
astro-ph/0601561
1
0601
2006-01-24T22:16:50
Color, Structure, and Star Formation History of Dwarf Galaxies over the last ~3 Gyr with GEMS and SDSS
[ "astro-ph" ]
We present a study of the colors, structural properties, and star formation histories for a sample of ~1600 dwarfs over look-back times of ~3 Gyr (z=0.002-0.25). The sample consists of 401 distant dwarfs drawn from the Galaxy Evolution from Morphologies and SEDs (GEMS) survey, which provides high resolution Hubble Space Telescope (HST) Advanced Camera for Surveys (ACS) images and accurate redshifts, and of 1291 dwarfs at 10-90 Mpc compiled from the Sloan Digitized Sky Survey (SDSS). The sample is complete down to an effective surface brightness of 22 mag arcsec^-2 in z and includes dwarfs with M_g=-18.5 to -14 mag. Rest-frame luminosities in Johnson UBV and SDSS ugr filters are provided by the COMBO-17 survey and structural parameters have been determined by S\'ersic fits. We find that the GEMS dwarfs are bluer than the SDSS dwarfs by ~0.13 mag in g-r, which is consistent with the color evolution over ~2 Gyr of star formation histories involving moderate starbursts and long periods of continuous star formation. The full color range of the samples cannot be reproduced by single starbursts of different masses or long periods of continuous star formation alone. Furthermore, an estimate of the mechanical luminosities needed for the gas in the GEMS dwarfs to be completely removed from the galaxies shows that a significant number of low luminosity dwarfs are susceptible to such a complete gas loss, if they would experience a starburst. On the other hand, a large fraction of more luminous dwarfs is likely to retain their gas. We also estimate the star formation rates per unit area for the GEMS dwarfs and find good agreement with the values for local dwarfs.
astro-ph
astro-ph
Color, Structure, and Star Formation History of Dwarf Galaxies over the last ∼ 3 Gyr with GEMS and SDSS Fabio D. Barazza, Shardha Jogee Department of Astronomy, University of Texas at Austin, 1 University Station C1400, Austin, TX 78712-0259, USA [email protected], [email protected] Hans-Walter Rix, Marco Barden, Eric F. Bell Max-Planck Institute for Astronomy, Koenigstuhl 17, D-69117 Heidelberg, Germany John A. R. Caldwell McDonald Observatory, University of Texas, Fort Davis, TX 79734, USA Daniel H. McIntosh Department of Astronomy, University of Massachusetts, Amherst, MA 01003, U.S.A Max-Planck Institute for Astronomy, Koenigstuhl 17, D-69117 Heidelberg, Germany Klaus Meisenheimer Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218, USA Chien Y. Peng and Christian Wolf Department of Physics, Denys Wilkinson Bldg., University of Oxford, Keble Road, Oxford OX1 3RH, U.K. ABSTRACT We present a study of the colors, structural properties, and star formation histories for a sample of ∼ 1600 dwarfs over look-back times of ∼ 3 Gyr (z = 0.002 − 0.25). The sample consists of 401 distant dwarfs drawn from the Galaxy Evolution from Morphologies and SEDs (GEMS) survey, which provides high – 2 – resolution Hubble Space Telescope (HST) Advanced Camera for Surveys (ACS) images and accurate redshifts, and of 1291 dwarfs at 10–90 Mpc compiled from the Sloan Digitized Sky Survey (SDSS). The sample is complete down to an effective surface brightness of 22 mag arcsec−2 in z and includes dwarfs with Mg = −18.5 to −14 mag. Rest-frame luminosities in Johnson UBV and SDSS ugr filters are provided by the COMBO-17 survey and structural parameters have been determined by S´ersic fits. We find that the GEMS dwarfs are bluer than the SDSS dwarfs by ∼ 0.13 mag in g − r, which is consistent with the color evolution over ∼ 2 Gyr of star formation histories involving moderate starbursts and long periods of continuous star formation. The full color range of the samples cannot be reproduced by single starbursts of different masses or long periods of continuous star formation alone. Furthermore, an estimate of the mechanical luminosities needed for the gas in the GEMS dwarfs to be completely removed from the galaxies shows that a significant number of low luminosity dwarfs are susceptible to such a complete gas loss, if they would experience a starburst. On the other hand, a large fraction of more luminous dwarfs is likely to retain their gas. We also estimate the star formation rates per unit area for the GEMS dwarfs and find good agreement with the values for local dwarfs. Subject headings: 1. Introduction Until recently the study of dwarf galaxies has been primarily concentrated either on clusters, where a large number of dwarfs can be observed within a relatively small area of sky (e.g., Binggeli & Cameron 1991; Trentham 1998; Drinkwater, Gregg, Holman, & Brown 2001), or on the Local Group, where dwarfs can be studied in great detail due to their proximity (for recent reviews see Mateo 1998; van den Bergh 2000; Grebel 2000). Dwarfs outside of these regimes are numerous, but widespread. Many of them are associated with giant galaxies, forming galaxy groups (Karachentsev 2005), which to some extent have been studied photometrically (Bremnes, Binggeli, & Prugniel 1998, 1999, 2000; Jerjen, Binggeli, & Freeman 2000) and kinematically (Bottinelli, Gouguenheim, Fouque, & Paturel 1990; Karachentsev, Makarov, & Huchtmeier 1999). Studies of dwarfs between clusters and groups are rather rare and are mainly composed of small samples, which have been selected more or less randomly (Makarova 1999; Barazza, Binggeli, & Prugniel 2001; Parodi, Barazza, & Binggeli 2002) or examine specific types of dwarfs, e.g., blue compact dwarfs (BCDs; Noeske, Papaderos, Cair´os, & Fricke 2003; Gil de Paz, Madore, & Pevunova 2003). – 3 – The evolution of dwarfs is a complex problem, where evolutionary paths may depend on a variety of external and internal factors. Our knowledge of the local volume (< 8 Mpc) has deepened, in particular due to strong efforts in determining distances to many nearby galaxies (Karachentsev et al. 2003, and references therein). However, it is still unclear what governs the evolution of dwarfs in low density regions and how the different morphological types form. Progress has in part been hampered by the fact that many early studies are based on small samples, suffer from small number statistics and had systematic selection biases. Headway toward constraining the evolutionary history of dwarf galaxies and identifying the fundamental physical processes involved requires, as a first step, the study of large, statistically significant samples of dwarfs, selected over a large region of the sky without systematic biases. In a cosmological context, empirical constraints on dwarf galaxies at different look-back times are essential for testing hierarchical models of galaxy formation (White & Frenk 1991; Steinmetz & Navarro 2002). One such aspect is the possibility that star formation in low mass dark matter halos gets suppressed due to a number of reasons, including the ejection of gas out of a shallow potential well by early, possibly primordial, episodes of star formation (Silk 2003; Burkert 2004). Many questions remain unresolved in simulations. Over what epochs does the suppres- sion happen: is this a primordial process only, or do dwarfs present a few Gyrs ago also exhibit signs of blowout? What types of masses and star formation histories do dwarf galax- ies exhibit empirically, and with such empirically established star formation histories, are they expected to retain their gas? What type of descendants would dwarfs at these epochs yield, if star formation was turned off? In this paper, we present a study of the properties of dwarf galaxies over the last 3 Gyr (z = 0.002 − 0.25)1 drawn from GEMS (Galaxy Evolution from Morphologies and SEDs, Rix et al. 2004) and the Sloan Digitized Sky Survey (SDSS, Abazajian et al. 2004). Our initial sample consists of 988 dwarfs from the GEMS survey in the redshift range z ∼ 0.09 − 0.25 (corresponding to look-back times Tback of 1 to 3 Gyr), and a comparison local sample of 2847 dwarfs with z < 0.02 (Tback ≈ 0.3 Gyr) from SDSS. In this paper, we concentrate on the colors, star formation rates (SFR), star formation histories, and structural properties of dwarf galaxies. In a follow-up paper (Barazza et al. in prep.) we will present a detailed morphological analysis of the dwarf sample from GEMS and also take into account their environment. Throughout the paper we will refer to the dwarf samples from GEMS and SDSS just as dwarf s without specifying classes (e.g., dE, Im etc.). However, we can assume 1We assume in this paper a flat cosmology with ΩM = 1 − ΩΛ = 0.3 and H0=70 km s−1 Mpc−1. – 4 – that most of them are Sm, Im, or BCD in the classification scheme of Sandage & Binggeli (1984), which was also confirmed by a rough visual inspection. This assumptions is based on the fact that both surveys particularly cover low density environments, where late type dwarfs dominate (Karachentsev et al. 2004). The rest of the paper is organized as follows: in § 2 we give a brief description of the GEMS survey and explain the sample selections; in § 3 follows the presentation of the comparisons performed and the resultant implications as well as a discussion of the findings; finally the summary and conclusions are given in § 4. 2. Observations, Sample Selection, and Analysis 2.1. GEMS survey GEMS is a two-band (F606W and F850LP) Hubble Space Telescope (HST ) large-area (800 arcmin2) survey using the Advanced Camera for Surveys (ACS), with accurate redshifts from Combo-17 (Wolf et al. 2004). The principal aim of the GEMS survey is the study of the evolution of galaxies out to Tback ∼ 8 Gyr (z ∼ 1.0), using morphologies and structural parameters as well as spectral energy distributions (SEDs). The morphological information is provided by a large-area (800 arcmin2) two-filter (V and z) imaging survey with ACS on HST . The GEMS field is centered on the Chandra Deep Field-South and reaches depths of = 28.3 (5σ) AB mag in F606W and = 27.1 (5σ) AB mag in F850LP for compact sources. The actual GEMS survey provides high-resolution (∼ 0.′′05 corresponding to 165 pc at z ∼ 0.2) ACS images for ∼ 8300 galaxies with accurate redshifts (δz/(1+z) ∼ 0.02 down to RV ega < 23 mag)2. Details of the data reduction and galaxy catalog construction are given in Caldwell et al. 2006, in prep. The dwarf sample drawn from the GEMS survey (see § 2.2) has a median redshift of z = 0.15. Therefore, the structural parameters from the GEMS z images correspond to a rest-frame filter between SDSS i and SDSS z. This has to be kept in mind and is further addressed in the discussion of the results presented below. On the other hand, the COMBO- 17 survey provides absolute magnitudes in rest-frames UBV and SDSS ugr, and can be compared directly with the local dwarfs. 2Henceforth, all magnitudes are given in the VEGA system – 5 – 2.2. The GEMS dwarf sample We identified and extracted dwarfs from the GEMS survey by applying an absolute magnitude cut of Mg > −18.5 mag. This limit is motivated by two findings. First the luminosity functions of different clusters exhibit small dips around MB ≈ −18 mag (e.g. Trentham & Hodgkin 2002; Mobasher et al. 2003; Pracy et al. 2004), which might indicate the transition between giants and dwarfs. Secondly, the typical luminosity, which separates Sd from Sm galaxies is around MB = −18 mag (Sandage, Binggeli, & Tammann 1985) and can be considered as the transition luminosity between disk galaxies massive enough to form spiral arms and disks with more irregular structures, commonly referred to as dwarfs. Thus, the applied cut also has morphological implications. The GEMS sample is complete down to R = 24 mag, which roughly corresponds to a surface brightness in z of µ ≈ 23.5 mag arcsec−2. However, we limit our GEMS dwarf sample to a effective surface brightness (µe) in z of µe < 22 mag arcsec−2, since this is the completeness limit of the SDSS sample (see § 2.3). We correct for the effect of cosmological surface brightness dimming using the factor (1 + z)4. Finally, we select the redshift range z = 0.09 − 0.25 for the GEMS dwarf sample. The lower limit corresponds to three times the expected error in redshift for our dwarf sample (δz/(1+z) ≈ 0.03), which is slightly larger than the average error of the GEMS sample, due to the lower luminosity of the dwarfs. The selection of the upper redshift limit was guided by the goal to target dwarfs over a large range of look-back times and potentially different evolutionary stages, while ensuring that the sample at the higher redshift end is not dominated by only bright galaxies, but includes a fair number of dwarfs in the range Mg = −14 to −18 mag. These selection conditions yield a sample of 401 objects. Our goal is to compare the GEMS sample to a sample from the NYU Value-Added low-redshift Galaxy Catalog (NYU-VAGC, Blanton et al. 2005a) of the SDSS, whose com- pleteness limits are defined with respect to effective surface brightness. We therefore fitted the GEMS dwarfs with a S´ersic model (S´ersic 1968) using GALFIT (Peng et al. 2002), which provides half-light radii and S´ersic indices for all objects. The fits were performed on the z-band images and we subsequently limited our final sample to µe ≤ 22 mag arcsec−2. In view of the completeness limit of the GEMS survey (see above), our final GEMS sample is complete down to µe = 22 mag arcsec−2 in z for objects with z < 0.25. The redshift distribution of the final sample, which consists of 401 objects, is shown in Figure 1. We note that the volume covered by this sample is rather small (∼ 17000 Mpc3) and the sample may suffer from cosmic variance and not be fully representative of the whole dwarf galaxy population at look-back times around 2 Gyr. Nonetheless, the advantage of this sample is that it allows us to study HST-based structural parameters of a complete sample of dwarfs at earlier look-back times (∼ 3 Gyr) than has been possible to date. – 6 – 2.3. SDSS low redshift dwarf sample SDSS is acquiring ugriz CCD imaging of 104 deg2 of the northern Galactic sky and selecting 106 targets for spectroscopy, most of them galaxies with r < 17.77 mag (Abazajian et al. 2004). The local dwarf sample is drawn from the low-redshift catalog of the NYU-VAGC (Blanton et al. 2005a), which is based on the second data release of SDSS.3. This low-redshift catalog consists of 28089 galaxies with distances of 10–200 Mpc (0.0033 < z < 0.05), which have been determined by correcting for peculiar velocities. The catalog provides, among other properties, rest-frame absolute magnitudes as well as photometric parameters from a S´ersic fit in the SDSS ugriz filters. The NYU-VAGC SDSS sample of dwarfs was selected using the limits Mg > −18.5 mag and z < 0.02. Finally, we again applied a cut at an effective surface brightness of 22 mag arcsec−2 in z, which corresponds to the completeness limit of the NYU-VAGC, given with respect to the r-band (Blanton et al. 2005a). The final sample of local dwarfs consists of 1291 objects and is complete down to our surface brightness limit. The luminosity distributions of the GEMS and SDSS samples are shown in Figure 2. The distributions are quite different in the sense that the fraction of low luminosity dwarfs (Mg > −16 mag) is larger in the GEMS sample. On the other hand, there is no large difference between the median values, which are −16.51 mag and −16.95 mag for the GEMS and SDSS samples, respectively. 3. Results and Discussion 3.1. Global colors The Combo-17 survey provides rest-frame magnitudes in the SDSS ugr bands, which allows us to directly compare the global colors of the GEMS and SDSS dwarf samples. The resulting color magnitude diagrams for both samples are shown in Figure 3a. The difference in global colors between the two samples is apparent in the histograms shown in Figure 3b. The median colors are 0.57 and 0.70 for GEMS and SDSS, respectively. A KS-test yields a probability of ∼ 2 × 10−41 that the two color distributions stem from the same parent distribution. In addition, there is a population of low luminosity, very blue GEMS dwarfs in Figure 3a, with hardly any local SDSS dwarf counterparts. We refer to these objects as low luminosity blue dwarfs (LLBDs) and include all objects with MB > −16.1 mag and B − V < 0.26 3The used data are therefore not affected by a recently described error in model magnitudes of extended objects (Strateva et al. 2005) – 7 – mag to this group. This leads to 48 LLBDs in the GEMS sample in the redshift range z = 0.09 − 0.23 (∼ 12% of the sample) and 8 LLBDs in SDSS (∼ 0.6% of the sample). The magnitudes and colors of the LLBDs are consistent with a recent intermediate- to low-mass starburst. We note, however, that the LLBDs are significantly bluer (B − V ∼ −0.1 to 0.26 mag) than typical blue compact dwarfs (B − V ≈ 0.45, Cair´os et al. 2001). On the other hand, there are two well studied dwarfs with luminosities and colors similar to the LLBDs: I Zw 18 (Izotov & Thuan 2004; Ostlin & Mouhcine 2005) and SBS 1415+437 (Thuan et al. 1999; Aloisi et al. 2005). However, we have to emphasize that the number of LLBDs in GEMS might be strongly overestimated. Some of these objects exhibit abnormally high IR emission, which indicates that they might in fact be star forming galaxies at much higher redshifts. Furthermore, 10 LLBDs exhibit a second smaller peak in their redshift probability distributions in COMBO-17, which typically occurs at z & 1. This caveat has to be kept in mind, whenever we discuss LLBDs in the following sections. Finally, the distribution of the SDSS galaxies defines a red sequence of dwarf galaxies, which has already been shown in the study of Blanton et al. (2005c). These red galaxies have been identified to be early type dwarfs, predominantly dwarf ellipticals. It is not clear, whether the GEMS dwarfs also exhibit a red sequence, probably because the number of objects is too small. 3.2. Structural parameters from S´ersic fits We determined structural parameters such as half light radii (re) and effective surface brightnesses (µe) by fitting a S´ersic model to the light distributions of the z-band images for both samples. The NYU-VAGC does provide such parameters for all objects. However, these parameters deviate systematically from the ones for the GEMS sample, which were determined using GALFIT. It seems that these differences have been introduced by the different fitting procedures and weighting schemes (see Appendix). Hence, we repeated the S´ersic fits to the SDSS objects using GALFIT, making sure that both samples are fitted using exactly the same procedure, software, and weighting scheme. In Figure 4 we show the histograms of µe in the z-band for both samples. The effective surface brightness is defined as the mean surface brightness within re. The two distributions agree quite well and the corresponding median values are very similar (21.10 mag arcsec−2 for GEMS and 21.21 mag arcsec−2 for SDSS). In Figure 5a we plot log(re) versus Mg for both samples and Figure 5b shows the histograms for log(re). The distributions overlap quite well, with the majority of GEMS dwarfs having comparable re to SDSS dwarfs. There is however a tail of low re values – 8 – exhibited by a small number of GEMS dwarfs. The median re for GEMS is 892 pc and for SDSS 1157 pc. A small contribution to this difference stems from the fact that, due to the redshift, the re for the GEMS dwarfs have been measured in a slightly bluer band, which is actually closer to the i-band than to the z-band for the median redshift. Therefore, we also fitted S´ersic models to the SDSS i-band images. The corresponding median value is 1125 pc. Thus, the difference in median values remains and might in part be caused by the red sequence dwarfs in SDSS (Figure 3a), which are much less prominent in GEMS. These objects are rather bright and therefore have rather large re. On the other hand, there is a significant population of Mg > −16 mag objects with log(re) below –0.5 in GEMS with almost no counterparts in SDSS (Figure 5a). These are the same objects, which make up the blue low luminosity group pointed out in § 3.1 (Figure 6). Finally, in Figure 7a the S´ersic index n is plotted versus Mg. For n = 1 the S´ersic model is equivalent to an exponential model and for n = 4 it is equivalent to a de Vaucouleurs model. The two distributions are quite similar, which is also exhibited in the histograms in Figure 7b. The distributions strongly peak around n = 1 (median value for GEMS is nmed = 1.32 and for SDSS nmed = 1.25), which indicates that the surface brightness profiles of most of the dwarfs are close to an exponential model. In fact, an exponential model is commonly found to best characterize the surface brightness profiles of late-type dwarfs (Makarova 1999; Barazza, Binggeli, & Prugniel 2001; Parodi, Barazza, & Binggeli 2002). On the other hand, for early-type dwarfs models with higher S´ersic indices seem to be more appropriate (Graham & Guzm´an 2003; Barazza et al. 2003). This provides some evidence that the GEMS and the SDSS samples are mainly composed of late-type dwarfs. There is also a group of low luminosity dwarfs in GEMS with quite high n. These are mainly the LLBDs, already identified in the previous sections. Hence, while the majority of GEMS and SDSS dwarfs have S´ersic index n < 2, the LLBDs are more compact and have therefore higher n. 3.3. Star formation histories of GEMS dwarfs In order to study the colors and star formation histories of the GEMS and SDSS dwarf samples, we compare their U −B and B −V colors to model tracks constructed using Starburst 99 (Leitherer et al. 1999; V´azquez & Leitherer 2005). We transform the SDSS colors to the Johnson colors using the following equations valid for the galaxy type Im: (U −B) = (u − g) − 0.99 and (B −V ) = (g − r) + 0.07 (Fukugita et al. 1995). Since we did not derive morphological classes for our samples, we cannot transform the colors accord- ing to galaxy types. This introduces some uncertainty to the Johnson colors of the SDSS – 9 – galaxies. However, as shown by the distribution of S´ersic indices (§ 3.2), both samples are strongly dominated by late-type dwarfs (galaxy types Sm, Im). In addition, we are able to roughly identify early type dwarfs according to their colors and have therefore some control of the uncertainties. Assuming that most galaxies with g − r & 0.85 are early-types, the color transformation discussed above adds an error to their U −B color as large as 0.36. In addition, there is another error source introduced by a red leak to the u filter and a bias in the sky subtraction reported on the SDSS web pages.4 Both uncertainties will predominantly affect galaxies with B −V & 0.6 and, in particular, affect their U −B color, causing it to be too red. Therefor, the star formation models are not expected to be able to reproduce the corresponding U −B, B −V color combinations. Thus, we will exclude all SDSS galaxies with U −B > 0.3 (see Figure 8) from the following discussion (242 objects, ∼ 19% of the SDSS sample), since their U −B colors are too uncertain. The line U −B = 0.3 is drawn on Figure 8 and 9 for reference. In Figure 8 the U −B versus B −V color-color plot is shown. A general difference between the two samples is again obvious. In order to understand, whether this difference could be linked to the evolution of the galaxies over ∼ 2 Gyr (corresponding to the look-back time of the GEMS sample), we compare the colors to different Starburst 99 model tracks. All models shown are based on a Kroupa Initial Mass Function (IMF, Kroupa 2001). The solid line in Figure 8 represents a model with a constant star formation rate (SFR) of 0.03 M⊙ yr−1 and a metallicity of Z = 0.004. These are typical values for field late type dwarfs (van Zee 2001; Hunter & Elmegreen 2004). The dashed line shows the color evolution of a Single Stellar Population (SSP) with a mass of 3 × 108 M⊙ and a metallicity of Z = 0.0004 formed in a single starburst. The dotted line represents the same SSP, but with a metallicity of Z = 0.008. The solid dots mark the following time steps (from left to right): 0.1, 0.5, 1.0, 2.0, 4.0, 8.0, 15.0 Gyr. All three models seem to be able to account for the colors of specific subgroups in Figure 8. However, they are most likely too simple, in order to represent the star formation histories of these dwarfs. The model with constant star formation does reproduce the colors of the bluest objects in both samples, but its luminosity remains too low (MB ≈ −12 mag) and we would have to apply an unreasonably high SFR (> 0.5M⊙ yr−1) to reach the observed luminosities. The single burst models are able to cover the colors of the redder objects in Figure 8 and they should be representative of galaxies, which do not form stars at the present time, i.e. of early type dwarfs. However, the stellar populations of early type dwarfs in the Local Group, the only such objects which can be studied in enough detail, are known to 4http://www.sdss.org/dr2/start/aboutdr2.html#imcaveat – 10 – be more complex than a SSP (Grebel 1998; Ikuta & Arimoto 2002; Tolstoy et al. 2003). It is therefore reasonable to assume that the star formation histories of dwarfs constitute a mixture of multiple bursts and periods of constant star formation, probably with different SFRs. Figure 8 suggests that the combination of different modes of star formation can reproduce the observed colors. In Figure 9 we plot tracks of models, which combine a single starburst with continuous star formation. The onset of the burst occurs at different times: together with the star formation at the earliest time step (solid line), 0.9 Gyr after the beginning of the continuous star formation (short dashed line), and 3.9 Gyr after the beginning of the continuous star formation (dotted line). In addition, we add a model with an exponentially decreasing SFR (long dashed line). These models are still just an approximation to the real star formation histories. For instance, the starburst component of the star formation histories could also be represented by several minor bursts, instead of one larger burst. However, the models reproduce the required luminosity range (−18 . MB . −14 mag between roughly 4–15 Gyr) and they cover the observed colors. The most important aspect of these model tracks is that they all exhibit a certain period of time, in which the U −B color remains more or less constant, but the B −V color changes significantly. This is in qualitative agreement with the B −V color difference between the samples shown in Figures 8 and 9, and also with the different distributions in Figure 3b. The models also show that an age difference of ∼ 2 Gyr is enough to increase the B −V color by ∼ 0.1, which is of the same order as observed. In addition, the color changes more rapidly at earlier times than at later times. This property is mostly caused by the burst component, whereas the passive evolution after the burst changes the colors only weakly. Interestingly, this characteristic is also shown by the observations. The median B −V color difference between the samples decreases with increasing U −B color. For the color bin −0.5 < U −B < −0.3 the difference is 0.16 and then decreases along the distribution reaching 0.01 for the bin 0.1 < U −B < 0.3. This indicates that the sequence of galaxies in the color- color plot is mainly due to the age of the most recent episode of star formation. The bluest galaxies are those having experienced a recent episode of star formation, while the redder dwarfs are older remnants. The luminosity appears to be less important for this relation, which is also indicated by the shallow slope in the color magnitude diagram (Figure 3a) and its large scatter. We can therefore conclude that the range of global colors shown by GEMS and SDSS dwarfs (over look-back times of 3 Gyr) is consistent with a star formation history involving multiple bursts of star formation, possibly combined with intermediate periods of relatively constant SFRs. Also a model representing an exponentially decreasing SFR is able to re- – 11 – produce the observed colors. In particular, we note that a single burst model or a history of constant star formation alone, cannot reproduce the full color range. We also find that the bluer colors typical of GEMS dwarfs (present at z = 0.09 − 0.25 or look-back times of 1–3 Gyr) can evolve by ∼ 0.1 mag into the redder colors of SDSS dwarfs (present at z < 0.02) over ∼ 2 Gyr. 3.4. Feedback from star formation and requirements for blowaway in dwarfs Empirical constraints on the luminosity, star formation histories, and star formation feedback in dwarf galaxies at different look-back times can provide useful constraints for hierarchical Lambda-CDM models of galaxy formation and, in particular, on the issue of feedback and blowout of gas via star formation in low mass galaxies. In a first step we, therefore, estimate the SFRs of the GEMS dwarfs. The COMBO-17 survey provides the rest-frame luminosities in a synthetic UV band centered on the 2800A line. For the redshift range of our GEMS sample the luminosities in this band are based on extrapolations beyond the filter set used in COMBO-17. For the luminosities in this UV band, we can derive the fluxes in the 2800A line, L2800. In order to estimate the SFRs we then use the following equation (Kennicutt 1998): SF R [M⊙yr−1] = 3.66 × 10−40L2800 [ergs s−1 λ−1] This equation applies to galaxies with continuous star formation over the last 108 years, which is likely the case for a majority of our dwarfs. For objects, which experienced a starburst within the last 108 years, the SFRs obtained in this way will underestimate the actual SFR. In Figure 10 we plot the normalized SFR versus MB. For the normalization we use the isophotal area provided by Sextractor. The range of normalized SFRs we obtained by this rough estimate agrees very well with the values found for a sample of nearby field dwarfs presented by Hunter & Elmegreen (2004). Figure 10 indicates that the LLBDs have relatively high SFRs per unit area. In order to obtain an estimate of the energies needed for an ejection of gas via starbursts in dwarf galaxies we consider the model of Mac Low & Ferrara (1999). The study estimates the impact of repeated supernovae on the interstellar medium. Two specific scenarios are distinguished: blowout, in which gas is expelled from the disk, but will in part be reaccreted at later epochs, and blowaway, in which the gas becomes completely unbound and is lost to the dwarf. The latter scenario is the one we consider for our estimate. The mechanical luminosity needed for a blowaway to occur depends on the visible mass of the dwarf and the effective sound speed, which in turn is related to the axis ratio (Mac Low & Ferrara 1999). – 12 – The corresponding equations are: L38 > 8 × 10−2M 7−6α vis,7 (cid:18) φ ω0(cid:19)6 c−10 10 h and ǫ = 0.43 φ ω0 M 1−α vis,7c−2 10 L38 corresponds to the logarithm of the mechanical luminosity in units of 1038ergs s−1, Mvis,7 is the visible mass in units of 107M⊙, α = 0.338 is a constant, φ is the dark-to-visible mass ratio given as φ ⋍ 34.7M −0.29 vis,7 , ω0 = 3 is a constant, c10 is the effective sound speed in units of 10 km s−1, h = H0/100, where H0 is the Hubble constant, and ǫ is the ellipticity. We estimate the visible mass of our galaxies using the V -band luminosity. The axis ratios are provided by the S´ersic fits. It is clear that more luminous galaxies need a higher mechanical luminosity for complete blowaway of the gas. This is due to the fact that we used the luminosity to derive the mass, which was then used to determine the mechanical luminosity. For a given MV , there is a range in mechanical luminosities for blowaway due to the range in ellipticities. Note, however, that the mechanical luminosities obtained in this way have to be considered as lower limits, since we most likely underestimate the mass and use apparent ellipticities. The estimates of the mechanical luminosities needed for a complete blowaway of the gas in our dwarfs (L38,BA) is plotted versus the mechanical luminosities inferred from the SFRs (L38,SF R) in Figure 11. The solid line corresponds to L38,SF R = L38,BA. Obviously, no significant gas loss would occur in these objects, if their SFRs would never exceed the ones inferred from their observed near-UV luminosities. Next, using the star formation histories and starburst strengths that were found to reproduce the observed range in colors of GEMS and SDSS dwarf galaxies (see § 3.3 and Figure 9), we computed the maximum mechanical luminosities, which can be injected into the ISM by the relevant starbursts of different masses. The range in luminosities is illustrated on Figure 11 by three dashed lines corresponding to starburst masses of 3 × 108 M⊙, 1 × 108 M⊙, 3 × 107 M⊙, and 1 × 107 M⊙. We find that for their derived star formation histories, the luminous (MB = −18 to −16 mag) dwarfs are likely to retain their gas and avoid blowaway. However, there are a fair number of low luminosity dwarfs (MB = −14 to −16) that are susceptible to a complete blowaway of gas, if they were to experience a starburst. However, in practice, only a small fraction of these low luminosity dwarfs may be actually undergoing a starburst. Even though, we do not have any clear evidence that some dwarfs in our sample experience a starburst at the time of observation, we are also not able to rule this out. The derived mechanical luminosities stem from the SFRs, which have been determined assuming that the dwarfs had a constant SFR over the last 108 years. In addition, we used the near-UV – 13 – luminosities, which could be affected by dust. In view of these uncertainties, the derived SFRs have to be considered as lower limits. The LLBDs could actually be a population, for which the SFRs have been significantly underestimated, since their magnitudes and colors are consistent with a recent low to intermediate starburst (§ 3.1) and they are compact (Figure 6), such that they will likely have a high SFR per unit area. This is confirmed in Figure 10. 4. Summary and Conclusions We present a study of the colors, structural properties, and star formation histories in a sample of ∼ 1600 dwarfs present over the last 3 Gyr (z = 0.002 − 0.25). Our sample consists of 401 dwarfs over z = 0.09 − 0.25 (corresponding to Tback of 1 to 3 Gyr) from the GEMS survey and a comparison sample of 1291 dwarfs with z < 0.02 (Tback < 0.3 Gyr) from the SDSS. Our final sample is complete down to an effective surface brightness of 22 mag arcsec−2 in z and includes dwarfs with Mg = −18.5 to −14 mag. The main results are: The mean global color of GEMS dwarfs (g − r = 0.57 mag) is bluer than that of the SDSS dwarfs (g − r = 0.70 mag) by 0.13 mag. Using Starburst 99, we find that the full range of global colors shown by SDSS and GEMS dwarfs, over look-back times of 3 Gyr, is consistent with a star formation history involving bursts of star formation, combined with periods of relatively constant SFRs or exponentially decreasing SFRs. In particular, we note that a single burst model or a history of constant star formation alone, cannot reproduce the full color range. We also find that the bluer colors typical of GEMS dwarfs can evolve by ∼ 0.1 mag into the redder colors of SDSS dwarfs over ∼ 2 Gyr. We identify a population of low luminosity (MB = −16.1 to −14 mag), blue (B −V < 0.26 mag) dwarfs (LLBDs) in the GEMS sample, with hardly any counterparts among local SDSS dwarfs. The very blue colors of the LLBDs are comparable to those of systems, such as I Zw 18 and SBS 1415+437. Their magnitudes and colors are consistent with a recent intermediate-to-low mass starburst. However, we have to stress the caveat that a large fraction of these objects might in fact be star forming galaxies at much higher redshifts. This is indicated by an abnormally high IR emission of some of these objects and the fact, that 10 LLBDs (out of 48) exhibit a second peak around z ∼ 1 in their redshift probability distributions. We performed S´ersic fits to the GEMS and SDSS dwarfs using GALFIT making sure that both samples are fitted using the exact same procedure and weighting schemes. Our experience shows that this approach is essential for avoiding large spurious differences (see – 14 – Appendix). We find that ∼ 76% of GEMS dwarfs and ∼ 81% of SDSS dwarfs have S´ersic n < 2. The majority of GEMS dwarfs have half light radii comparable to those of SDSS dwarfs. We estimate the SFR per unit area of GEMS dwarfs using the rest-frame UV (2800A) luminosity and find values in the range 5 × 10−3 to 5 × 10−1 M⊙ yr−1 kpc−2. We note that the LLBDs have magnitudes and colors consistent with a recent low-to-intermediate mass starburst, small half light radii, and resultant high SFRs per unit area. Finally, we estimate the mechanical luminosities needed for the gas in the GEMS dwarfs to become unbound and lost (blowaway). We then compare these to maximum mechanical power that would be injected by starbursts that are consistent with the derived star formation histories of these dwarfs. We find that the luminous (MB = −18 to −16 mag) dwarfs are likely to retain their gas and avoid blowaway. However, there are a fair number of low luminosity dwarfs (MB = −16 to −14 mag) that are susceptible to a complete blowaway of gas, if they were to experience a starburst. F.D.B. and S.J. acknowledge support from the National Aeronautics and Space Admin- istration (NASA) LTSA grant NAG5-13063 and from HST-GO-10395 and HST-GO-10428. E.F.B. was supported by the European Community's Human Potential Program under con- tract HPRN-CT-2002-00316 (SISCO). C.W. was supported by a PPARC Advanced Fellow- ship. D.H.M acknowledges support from the NASA LTSA Grant NAG5-13102. C.Y.P is grateful for support by the Institute Fellowship at STScI. Support for GEMS was provided by NASA through number GO-9500 from the Space Telescope Science Institute STScI, which is operated by the Association of Universities for Research in Astronomy, Inc. AURA, Inc., for NASA, under NAS5-26555. A. Systematic differences in S´ersic parameters derived using different fitting methods During the analysis of the structural properties of the two samples we realized that there are systematic differences between the S´ersic parameters provided by the NYU VAGC and the results from GALFIT (Peng et al. 2002) used for the GEMS sample. The differences are larger than the expected errors and are most likely caused by specifics of the two fitting methods applied. In order to confirm this conjecture, we randomly selected 250 galaxies from SDSS covering the entire luminosity range considered in this study. We fitted the surface brightness distributions of this subsample with a S´ersic model using GALFIT. Five objects could not be fitted by GALFIT. The effective radii (re) of the remaining 245 galaxies are compared – 15 – to the re provided by the NYU VAGC in Figure 12. Obviously, GALFIT is systematically measuring larger radii. The fitting method used for the NYU VAGC is described in Blanton et al. (2005b). There appear to be two major differences in the fitting procedure compared to GALFIT: (1) all pixels are weighted equally in NYU VAGC, whereas in GALFIT each pixel is weighted by the Poisson noise, (2) in the NYU VAGC a axissymmetric model is assumed, whereas in GALFIT the ellipticity and the position angle are included in the fitting process. The latter discrepancy probably makes the largest contribution to the differences observed, as suggested by Figure 13. The largest differences occur in the most inclined objects, although with a very large scatter. After fitting both samples using the same method and weighting scheme as in GEMS, the differences in re practically disappear (see § 3.2). The presented comparison does not intend to judge the quality or correctness of the two approaches, but it shows that one has to be very cautious by combining model parameters stemming from different fitting methods. Abazajian, K., et al. 2004, AJ, 128, 502 REFERENCES Aloisi, A., van der Marel, R. P., Mack, J., Leitherer, C., Sirianni, M., & Tosi, M. 2005, ApJ, 631, L45 Barazza, F. D., Binggeli, B., & Prugniel, P. 2001, A&A, 373, 12 Barazza, F. D., Binggeli, B., & Jerjen, H. 2003, A&A, 407, 121 Binggeli, B. & Cameron, L. M. 1991, A&A, 252, 27 Blanton, M. R., et al. 2005a, AJ, 129, 2562 Blanton, M. R., Eisenstein, D., Hogg, D. W., Schlegel, D. J., & Brinkmann, J. 2005b, ApJ, 629, 143 Blanton, M. R., Lupton, R. H., Schlegel, D. J., Strauss, M. A., Brinkmann, J., Fukugita, M., & Loveday, J. 2005c, ApJ, 631, 208 Bottinelli, L., Gouguenheim, L., Fouque, P., & Paturel, G. 1990, A&AS, 82, 391 Bremnes, T., Binggeli, B., & Prugniel, P. 1998, A&AS, 129, 313 Bremnes, T., Binggeli, B., & Prugniel, P. 1999, A&AS, 137, 337 – 16 – Bremnes, T., Binggeli, B., & Prugniel, P. 2000, A&AS, 141, 211 Bruzual A., G., & Charlot, S. 1993, ApJ, 405, 538 Burkert, A. 2004, ASP Conf. Ser. 322: The Formation and Evolution of Massive Young Star Clusters, 322, 489 Cair´os, L. M., Caon, N., V´ılchez, J. M., Gonz´alez-P´erez, J. N., & Munoz-Tun´on, C. 2001, ApJS, 136, 393 Drinkwater, M. J., Gregg, M. D., Holman, B. A., & Brown, M. J. I. 2001, MNRAS, 326, 1076 Fukugita, M., Shimasaku, K., & Ichikawa, T. 1995, PASP, 107, 945 Gil de Paz, A., Madore, B. F., & Pevunova, O. 2003, ApJS, 147, 29 Graham, A. W., & Guzm´an, R. 2003, AJ, 125, 2936 Grebel, E. K. 1998, Highlights in Astronomy, 11, 125 Grebel, E. K. 2000, ESA SP-445: Star Formation from the Small to the Large Scale, 87 Hunter, D. A. & Elmegreen, B. G. 2004, AJ, 128, 2170 Ikuta, C., & Arimoto, N. 2002, A&A, 391, 55 Izotov, Y. I., & Thuan, T. X. 2004, ApJ, 616, 768 Jerjen, H., Binggeli, B., & Freeman, K. C. 2000, AJ, 119, 593 Karachentsev, I. D., Makarov, D. I., & Huchtmeier, W. K. 1999, A&AS, 139, 97 Karachentsev, I. D., et al. 2003, A&A, 398, 479 Karachentsev, I. D., Karachentseva, V. E., Huchtmeier, W. K., & Makarov, D. I. 2004, AJ, 127, 2031 Karachentsev, I. D. 2005, AJ, 129, 178 Kennicutt, R. C. 1998, ARA&A, 36, 189 Kroupa, P. 2001, MNRAS, 322, 231 Leitherer, C., et al. 1999, ApJS, 123, 3 – 17 – Mac Low, M.-M., & Ferrara, A. 1999, ApJ, 513, 142 Makarova, L. 1999, A&AS, 139, 491 Mateo, M. L. 1998, ARA&A, 36, 435 Mobasher, B., et al. 2003, ApJ, 587, 605 Noeske, K. G., Papaderos, P., Cair´os, L. M., & Fricke, K. J. 2003, A&A, 410, 481 Ostlin, G., & Mouhcine, M. 2005, A&A, 433, 797 Parodi, B. R., Barazza, F. D., & Binggeli, B. 2002, A&A, 388, 29 Peng, C. Y., Ho, L. C., Impey, C. D., & Rix, H. 2002, AJ, 124, 266 Pracy, M. B., De Propris, R., Driver, S. P., Couch, W. J., & Nulsen, P. E. J. 2004, MNRAS, 352, 1135 Rix, H., et al. 2004, ApJS, 152, 163 Sandage, A. & Binggeli, B. 1984, AJ, 89, 919 Sandage, A., Binggeli, B., & Tammann, G. A. 1985, AJ, 90, 1759 Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525 S´ersic, J. L. 1968, Cordoba, Argentina: Observatorio Astronomico, 1968 Silk, J. 2003, MNRAS, 343, 249 Steinmetz, M., & Navarro, J. F. 2002, New Astronomy, 7, 155 Strateva, I. V., et al. 2005, AJ, 130, 1961 Thuan, T. X., Izotov, Y. I., & Foltz, C. B. 1999, ApJ, 525, 105 Tolstoy, E., Venn, K. A., Shetrone, M., Primas, F., Hill, V., Kaufer, A., & Szeifert, T. 2003, AJ, 125, 707 Trentham, N. 1998, MNRAS, 293, 71 Trentham, N. & Hodgkin, S. 2002, MNRAS, 333, 423 van den Bergh, S. 2000, The galaxies of the Local Group, by Sidney Van den Bergh. Published by Cambridge, UK: Cambridge University Press, 2000 Cambridge Astrophysics Series Series, vol no: 35, ISBN: 0521651816., – 18 – van Zee, L. 2001, AJ, 121, 2003 V´azquez, G. A., & Leitherer, C. 2005, ApJ, 621, 695 White, S. D. M., & Frenk, C. S. 1991, ApJ, 379, 52 Wolf, C., et al. 2004, A&A, 421, 913 This preprint was prepared with the AAS LATEX macros v5.2. – 19 – Fig. 1.- The number of objects per redshift bin of our final GEMS sample (401 objects). The upper axis shows the approximate look-back time. The median look-back time of the sample is ∼ 1.8 Gyr. The inhomogeneous distribution is caused by cosmic variance. – 20 – Fig. 2.- The g-band luminosity distribution of the GEMS and SDSS samples. The me- dian value for the GEMS dwarf sample is −16.51 mag and for SDSS −16.95 mag. The completeness limits of the GEMS sample at two redshifts are indicated. – 21 – Fig. 3.- a) The color magnitude diagram for the GEMS and SDSS samples. The AB mag- nitudes given in SDSS have been transformed to Vega magnitudes using the transformations given in Wolf et al. (2004). The error bar applies to both samples. The two dashed lines indicate the typical colors of early type dwarfs in the Local Group (upper line, Mateo 1998) and of BCDs (lower line, Gil de Paz, Madore, & Pevunova 2003). There are specific g − r values (e.g. at g − r = 0.3), around which the GEMS dwarfs seem to cluster, resulting in some distinct horizontal features. These stem from the template fitting process and are not real. The resulting gaps in the distribution are smaller than the errors and are therefore not affecting the results. b) The g − r color distributions for both samples. The median colors are 0.57 and 0.70 for GEMS and SDSS, respectively. – 22 – Fig. 4.- Histogram of the effective surface brightness distribution (µe) for the GEMS and SDSS samples. µe is the mean surface brightness in z within the effective (half light) radius, which was determined by a S´ersic fit to the z-band images. The median values are 21.10 mag arcsec−2 and 21.21 mag arcsec−2 for GEMS and SDSS, respectively. – 23 – Fig. 5.- a) The logarithm of the half light radius in kpc versus Mg. The half light radii have been determined from a S´ersic fit to the z-band images using GALFIT. A linear fit to the GEMS sample gives −0.20 × Mg − 3.33; the median half light radius is 892 pc. For the SDSS sample we obtain −0.18 × Mg − 2.98 and 1157 pc. b) Histograms of the logarithms of the half light radii for both samples. – 24 – Fig. 6.- The g − r color versus the logarithm of the half light radius in kpc. – 25 – Fig. 7.- a) The shape parameter n from the S´ersic fits versus Mg. A value of n = 1 corresponds to a exponential disk, whereas a value of n = 4 is equivalent to a de Vaucouleurs model. b) Histograms of n. The median values are 1.32 and 1.25 for GEMS and SDSS, respectively. – 26 – Fig. 8.- A B − V versus U − B color-color plot. Overplotted are three Starburst 99 model tracks, which are all computed using a Kroupa IMF. The black dots indicate the following time steps (from left to right): 0.1, 0.5, 1.0, 2.0, 4.0, 8.0, 15.0 Gyr. Solid line: constant star formation with SF R = 0.03 M⊙ yr−1 and Z = 0.004. Dashed line: single starburst with a mass of 3 × 108 M⊙ and Z = 0.0004. Dotted line: single starburst with a mass of 3 × 108 M⊙ and Z = 0.008. Objects above the dashed line have uncertain U −B colors and are excluded from the discussion. The error bar represents only the error from the measurement and does not include uncertainties stemming from the color transformation for SDSS. The arrow indicates the effect of dust on the colors (Schlegel et al. 1998). – 27 – Fig. 9.- The same as Figure 8. Three lines represent models, where a continuous star formation with a constant rate of SF R = 0.03 M⊙ yr−1 and a metallicity of Z = 0.004 has been combined with various single starbursts starting at different times. For these models a Kroupa IMF has been used. The fourth line represents a model with an exponentially decreasing star formation rate and a Salpeter IMF. Solid line: A single starburst with a mass of 3 × 108 M⊙ and Z = 0.0004 starts at 0.1 Gyr. The time steps are the same as in Figure 8. Short dashed line: A single starburst with a mass of 3 × 108 M⊙ and Z = 0.004 starts at 0.9 Gyr. The time steps are: 0.1, 0.5, 1.0, 1.1, 1.5, 2.0, 4.0, 8.0, 15.0. Dotted line: A single starburst with a mass of 5 × 108 M⊙ and Z = 0.02 starts at 3.9 Gyr. The time steps are: 0.1, 0.5, 1.0, 2.0, 4.0, 4.1, 4.5, 8.0, 15.0. Long dashed line: An exponentially decreasing SFR, using models from Bruzual & Charlot (1993). The time steps are: 0.1, 1.0, 3.0, 10.0 Gyr. The arrow indicates the effect of dust on the colors (Schlegel et al. 1998). – 28 – Fig. 10.- The normalized SFR versus MB. The SFRs have been estimated from the 2800A continuum fluxes (L2800) and using the equation SF R[M⊙yr−1] = 3.66 × 10−40L2800 [ergs s−1 λ−1] adopted from Kennicutt (1998). These SFRs have then been divided by the isophotal area provided by Sextractor. – 29 – Fig. 11.- Plot of the mechanical luminosity needed for a complete blowaway of the gas in dwarfs having a given MV and a range of apparent ellipticities versus the mechanical luminosities inferred from the SFRs. L38 corresponds to the logarithm of the mechanical luminosity in units of 1038ergs s−1. The four dashed lines mark the peak mechanical lu- minosities reached of starbursts with the indicated masses. The solid line corresponds to L38,SF R = L38,BA. – 30 – Fig. 12.- Histograms of re for a subsample of 245 galaxies from SDSS. The values obtained by a S´ersic fit using GALFIT are compared to the values provided by the NYU VAGC. – 31 – Fig. 13.- The logarithm of the ratios of re measured by GALFIT and VAGC versus apparent ellipticity. The ellipticity stems from GALFIT.
astro-ph/0212268
1
0212
2002-12-11T18:47:17
A simplified model of the formation of structures in the dark matter, and a background of very long gravitational waves
[ "astro-ph" ]
Collapse of the rotating spheroid is approximated by a system of ordinary differential equations describing its dynamics. The gravitational potential is approximated by the one of the iniform Maclaurin spheroid. Developement of gravitational instability and collapse in the dark matter medium do not lead to any shock formation or radiation, but is characterized by non-collisional relaxation, which is accompanied by the mass and angular momentum losses. Phenomenological account of these processes is done in this model. Formation of the equilibrium configuration dynamics of collapse is investigated for several parameters, characterizing the configuration. A very long gravitational wave emission during the collapse is estimated, and their possible connection with the observed gravitational lenses is discussed.
astro-ph
astro-ph
A simplified model of the formation of structures in the dark matter, and a background of very long gravitational waves G.S. Bisnovatyi-Kogan ∗ Abstract Collapse of the rotating spheroid is approximated by a system of or- dinary differential equations describing its dynamics. The gravitational potential is approximated by the one of the iniform Maclaurin spheroid. Developement of gravitational instability and collapse in the dark matter medium do not lead to any shock formation or radiation, but is character- ized by non-collisional relaxation, which is accompanied by the mass and angular momentum losses. Phenomenological account of these processes is done in this model. Formation of the equilibrium configuration dynamics of collapse is investigated for several parameters, characterizing the con- figuration. A very long gravitational wave emission during the collapse is estimated, and their possible connection with the observed gravitational lenses is discussed. 1 Introduction A study of of a formation of dark matter objects in the universe is based on N-body simulations, which are very time consuming. In this situation a sim- plified approach may become useful, because it permits to investigate rapidly by analytical or by simple numerical calculations many different variants , and obtain crudly some new principally important features of the problem which could be lost or not visible during a long numerical work. The modern theory of a large scale structure is based on the ideas of Zel- dovich (1970) about a formation of strongly non-spherical structures during non-linear stages of a development of the gravitational instability, known as "Zeldovich's pancakes". The numerical simulations started by Doroshkevich et al. (1980) had been performed subsequently by many groups, revealing compli- cated structures of the new born objects (see e.g. Doroshkevich et al., 1999). Here we derive and solve equations for a simple model of a dynamical be- haviour of a compressible rotating spheroid in which a motion along both axies ∗Space Research Institute, Moscow, Russia, Profsoyuznaya 84/32, Moscow 117810, Russia; [email protected]; and Theoretical Astrophysics Center, Copenhagen, Denmark 1 takes place in the common gravitational field of a uniform spheroid, and un- der the action of the isotropic pressure (taken into account in the approximate non-differential way), and otherwise independently on each other. Due to non- interacting particles of the dark matter a strong compression does not lead to a formation of the shock wave and additional energy losses due to radiation at increasing of a temperature.. All losses are connected with a runaway particles, and relaxation consists only in the phase mixing leading effectively to a trans- formation of the kinetic energy of the ordered motion (collapse) into the kinetic energy of the chaotic motion, and to creation an effective pressure and thermal energy. In presence of the chaotic motion the pancake formed during the col- lapse has a finite minimal thickness, because particles cross the equatorial plane non-simultaneously, forming therefore the effective pressure, which "stops" the contraction. These effects may be described approximately in the hydrodynam- ical approach, in which there are no shock waves and the main transport process is an effective bulk viscosity, leading to relaxation and to a damping of the or- dered motion. This relaxation is based on the idea of a "violent relaxation" of Lynden-Bell (1967). The system of the ordinary differential equations is derived, which describes the dynamical behaviour of the compressible spheroid, where the relaxation and losses of the energy, mass, and angular momentum are taken into accound phenomenologically. In absence of any dissipation these equations describe a self-consistent conservative system, where gravitational and thermal pressure forces are present. Namely, the equilibrium solution of these equa- tion describes a Maclaurin spheriod, where the density is not prescribed but is found self-consistently, when the effective entropy is known. We consider here only non-relativistic dark matter with a relations between the density ρ, energy density E, pressure P , and specific entropy S as ρE = 3 2 P, P ∼ ρ5/3, E ∼ ρ2/3 at constant S. Solution of simple ordinary equations reveals some interesting features of the collapse of non-interacting matter, which had not been noticed earlier. The collapse of a spherioid, leading to the formation of a pancake may be continued by the fornation of a transient oblate figure, strongly elongated along the axis of a symmetry with a ratio of the axis c to the equatorial radius a : c/a ∼ 5. Formation of such transient figures, appearing only at low rotation, could be expected in central parts of the collapsing dark matter objects during initial stages of the collapse, when the influence of the outer non-uniform part of the cloud is still negligible. We have found, that at a realistic rate of the relaxation rate with a character- istic time equal to 3 local Jeans times, the oscillations do not damp completely, and about 1/10 of the initial amplitude of a velocity survives after at least 10 oscillationl periods. That means that the most massive dark matter objects may be still in the oscillatory state. These oscillations, as well as variable grav- itational fields of the collapsing dark matter objects could be important for the the interpretation of the observed picture of the cosmic microvave background (CMB) fluctuations. Very long gravitational waves emitted mainly during the 2 first stage of the pancake formation (see Thuan and Ostriker, 1974; Novikov, 1976) could be also important for the CMB fluctuations (Doroshkevich et al., 1977), as well as for the gravitational lensing of the distant objects. 2 Properties of a Maclaurin spheroid Let us consider a spheroid with semi-axies a = b > c x2 + y2 a2 + z 2 c2 = 1, (1) and rotating uniformly with an angular velocity Ω. Let us approximate the density of matter ρ in the spheroid as uniform. The mass m and total angular momentum M of the spheroid are connected with the angular velosity and semi- axies as (Landau, Lifshits, 1988) m = 4π 3 a2cρ, M = 2 5 ma2Ω. Introducing k = c/a < 1, we may express the gravitational energy Ug as (2) (3) Ug = − 3Gm2 5a arccos k √1 − k2 , what gives U = −3Gm2/5a for a sphere. The gravitational forces inside the uniform spheriod Fx, Fy, Fz are expressed via the grvitational potential φ as F = −∇φ, and are defined by the relations Fx = 2πGρx 1 − k2 (cid:20)k2 − Fz = − k arccos k √1 − k2 (cid:21), Fy = √1 − k2 (cid:21). k arccos k 4πGρz 1 − k2(cid:20)1 − y x Fx, (4) The gravitational potential with normalization φ → 0 at infinity, inside the spheroid is expressed as φ = − 1 2 (xFx + yFy + zFz) − 2πGρa2 k arccos k √1 − k2 . (5) During the process of relaxation a form of the spheroid changes between oblate and prolate ones. When c > a = b the following relations are valid instead of (3)-(5) 3Gm2 5a Archk √k2 − 1 , Ug = − k2 − 1(cid:20)k2 − 2πGρx Fx = − kArchk √k2 − 1(cid:21), Fy = y x Fx, 3 (6) (7) Fz = 4πGρz k2 − 1(cid:20)1 − kArchk √k2 − 1(cid:21). φ = − 1 2 (xFx + yFy + zFz) − 2πGρa2 kArchk √k2 − 1 . (8) Here the positive branch of Archk should be used. 3 Equations of motion For a real Maclaurin equilibrium noncompressible spheroid the oblateness k is connected with the angular velocity (or angular momentum) k(1 + 2k2) (1 − k2)3/2 arccos k − 1 − k2 = 3k2 Ω2 2πGρ = 25 6G M 2 m3 k a . (9) We consider a case of non-collisional particles when the density varies in time, and approximate the form and gravitational potential of the body by the form and gravitational potential of Maclaurin spheroid. Here pressure does not ne- sessary balance the garvitational and centrifugal forces. Consider a uniform spheroid with a constant mass and angular momentum, and with a total "ther- mal" energy of non-relativistic dark matter particles Eth ∼ V −2/3 ∼ (abc)−2/3 (V is a volume of a spheroid, a = b). Equations of motion for this spheroid are written for the radial accelerations at the equator and at the pole for semiaxes a, c a = 25 4 M 2 m2a3 + 3 2 Gm a2(1 − k2)(cid:20)k − c = −3 Gm a2(1 − k2)(cid:20)1 − arccos k √1 − k2(cid:21) + p1 − k2)(cid:21) + k arccos k 10Eth,in abc (cid:19)2/3 3am (cid:18) ainbincin abc (cid:19)2/3 3cm (cid:18) ainbincin . 10Eth,in , (10) (11) Here Eth,in, ain = bin, cin are the initial values of corresponding parameters. Equations (10),(11) describe the dynamics of the concervative system with a linear dependence of the velocity on the coordinates vx = ax a , , vy = by b , vz = cz c , and Lagrange function L = Ukin − Upot, Upot = Urot + Ug + Eth, M 2 5 ma2 , Eth = 4 (abc)2/3 , E = Eth,in(ainbincin)2/3 E Urot = (12) (13) The function Ug defined in (3), and the rotational energy from (13) should be used during the variation of L in the form 4 Ug = − 6Gm2 5(a + b) arccos k′ √1 − k′2 , The kinetic energy is equal to k′ = 2c a + b , Urot = 5 2 M 2 m(a2 + b2) . (14) Ukin = 1 2 a (cid:19)2 ρZV(cid:20)(cid:18) ax b !2 + by +(cid:18) cz c (cid:19)2(cid:21)dx dy dz = m 10 ( a2 + b2 + c2). (15) Equations of motion (10),(11) are the Lagrange equations with the Lagrange function (13)1 Solutions of these equations describe either pure oscillations, or the total disruption, depending on the initial conditions. For the prolate spheroid the following relations are valid instead of (10),(11),(14) a = 25 4 M 2 m2a3 − 3 2 Gm a2(k2 − 1)(cid:20)k − Archk √k2 − 1(cid:21) + pk2 − 1)(cid:21) + kArchk 10Eth,in abc (cid:19)2/3 3am (cid:18) ainbincin abc (cid:19)2/3 3cm (cid:18) ainbincin , 10Eth,in , (16) (17) , k′ = , Urot = 2c a + b 5 2 M 2 m(a2 + b2) . (18) c = 3 Gm a2(k2 − 1)(cid:20)1 − Ug = − 6Gm2 5(a + b) Archk′ √k′2 − 1 4 Equations of motion with dissipation In the reality there is a relaxation in the collisionless system, connected with a phase mixing which called "violent relaxation" (Lynden-Bell, 1967). This relaxation leads to a dissipation of the energy of the kinetic motion and increase of the chaotic (thermal) energy and pressure. As a result of this dissipation the kinetic motion will suffer from an effective drag force, which account is described phenomenologically by adding of the terms a τrel − and − c τrel (19) 1We consider here for simplicity that the relaxation is accompanied by the isotropization of the distribution function in the velociyu space. The violent relaxation has a non-collisional origin, and if the anizotropic instability has a lower increment the anizotropy in the velocity space may be preserved. In the limiting case of a pure anizotropic relaxation we have three (or two in our case) pressure and entropy functions, so that Eth = m(cid:16) Ex a2 + Ey b2 + Ez c2(cid:17) , Px = 2 3 ρ Ex a2 , Py = 2 3 ρ Ey b2 , Pz = 2 3 ρ Ez c2 . We should have the terms !0Ex/a3m and !0Ez/c3m instead of the last terms in (10),(11), respectively. A corresponding changes should be done also in the equations determining increase of entropies, and different losses. 5 in the right parts of equations (10),(11) respectively. We consider the nonrela- tivistic dark matter with the relation between the pressure P and the thermal energy Eth of the whole spheroid as Eth = 3 2 P m ρ . (20) Dissipation (19) leads to a heat production. Let us find a rate of a heat pro- duction leading to the growth of the entropy of the spheroid matter. Write equations of motion (10),(11) with account of (5),(13),(19) separately for a, b and c, allowing also the energy, mass and angular momentum losses ∂Ug ∂a − a ∂Urot ∂a − a ∂Eth ∂a − 1 5 m a2 τrel , 1 m 1 m 1 m d d dt(cid:18) m2 a2 dt m2 b2 dt(cid:18) m2 c2 10 (cid:19) = − a 5 ! = −b 5 (cid:19) = − c d ∂Ug ∂b − b ∂Urot ∂b − b ∂Eth ∂b − ∂Eth ∂c − ∂Ug ∂c − c ∂Urot ∂c − c τrel = 2παrelr abc 3Gm . 1 5 1 5 m b2 τrel , m c2 τrel , (21) (22) (23) (24) Note that in (21)-(23) the gravitational energy should be taken in the form (14). We have scaled here the relaxation time τrel by the Jeans characteristic time τJ = 2π ωJ = 2π √4πGm = 2πr abc 3Gm , (25) with a constant value of αrel. Collecting together (21)-(23) we find the energy balance relation in the form dUtot dt = ∂Ug ∂m dm dt + ∂Urot ∂m + dm dt ∂Urot ∂M dE dt −2 Utot = Ukin + Ug + Urot + Eth. ∂Eth ∂E dM dt + Ukin τrel − Ukin m dm dt , (26) The process of relaxation is accompanied also by the energy, mass, and angular momentum losses from the system. We suggest, that these losses take place only in non-stationary phases, so these rates are proportional to the kinetic energy Ukin, the characterictic times for mass, angular momentum and energy losses τml, τMl,τel are considerably greater then τrel. The equations describing different losses may be phenomenologically written as dUtot Ukin τml , dt = − mUkin Ugτml dm dt = , Ug < 0, 6 (27) (28) dM dt = M Ukin UgτMl . (29) Scaling all the characteristic times by the Jeans value, like in the case (24) of τrel, we have τel = αelτJ , τml = αmlτJ , τMl = αMlτJ , (30) with constant values of αi (i = el, ml, M l). Combining (28),(29), we find a relation M Min min(cid:19) =(cid:18) m α ml α M l , (31) where min and Min are the initial values of corresponding parameters. The function E determines the entropy of the matter in the spheroid, so that, using (13), we have dEth + P dV dt = 1 (abc)2/3 dE dt , (32) dt E Eth = (abc)2/3 , V = 4π 3 abc, P = 2 3 Eth V . Using (27)-(29) in (26) we obtain the equation for the entropy function E in presence of different losses in the form dE dt = (abc)2/3Ukin(cid:20)(cid:18) 2 τrel − 1 τel − 2 τml(cid:19) − Urot Ug (cid:18) 2 τMl − 1 τml(cid:19) + Ukin Ugτml(cid:21). (33) 5 Non-dimensional equations The equations, describing approximately the dynamics of the fomation of a sta- tionary dark matter object include equations of motion (10),(11) with adding terms from (19); energy equation (33) with account of (3),(13),(15); and equa- tions (27)-(29), describing the losses of the energy, mass and angular momen- tum. The mass losses are also taken into account in the equations of motion, written in the form (21)-(23). For obtaining a numerical solution of these equa- tions let us write them in non-dimensional variables. Introduce the following non-dimensional variables , c = c a0 , m = m m0 , M = M M0 , (34) , Eth = Eth U0 , E = E E 0 , τi = τi t0 . The process, with account of mass and angular momentum losses, is described by the following non-dimensional system of equations t = ρ = t t0 ρ ρ0 , a = , U = a a0 U U0 7 a = a m a(2 a2 + c2) 6τmlak + 25 4 M 2 m2a3 + 3 2 m a2 a1k − a τrel + 10 3am E (a2c)2/3 , c = c m a(2 a2 + c2) 6τmlak − 3 m(2 a2 + c2) dE dt = 10 + 25 12 M 2 m3a ak (cid:18) 2 dm dt = − a(2 a2 + c2) 6τmlak , E 10 3cm 2 + (a2c)2/3 , τml(cid:19) 1 τel − a(2 a2 + c2) 6mτmlak (cid:21), τrel − α ml α M l , Ω = 5 2 M ma2 , c τrel m a2 a2k − (a2c)2/3(cid:20)(cid:18) 2 τml(cid:19) + τMl − =(cid:18) m min(cid:19) 1 M Min df dt = Ω. k − ak 1 − k2 , a2k = 1 − k ak 1 − k2 Here ak = arccos k √1 − k2 , a1k = for the oblate spheriod a = b > c, and (35) (36) (37) (38) (39) (40) (41) ak = Archk √k2 − 1 , a1k = − k − ak k2 − 1 , a2k = − 1 − k ak k2 − 1 In both cases ak + a1k = k a2k. for the prolate spheriod with a = b < c. The rotation phase f is calculated in (39) for comparison of the rotation and oscillation processes. When the spheroid is close to the sphere k − 1 ≪ 1 we have the following expansions ak ≈ 7 6 − 4 3k − k2 6 1 3k3 , a1k ≈ − 1 1 k , a1k ≈ − k + 1 − + 1 3k3 , a2k ≈ 1 1 6 , a2k ≈ ak ≈ 1 3k2 at a = b > c, (42) k + 1 − k 6 at a = b < c. In the case of a very compressed spheroid (pancake) we have the following expansion ak = π 2 − k + π 4 k2 − 2 3 k3, a1k = − π 2 + 2k − 3π 4 k2, a2k = 1 − π 2 k + 2k2. (43) In equations (35)-(39) only non-dimentional variables are used, and "tilde" sign is omitted for simplicity. The non-dimensional relaxation times τi in these equations are written with account of (25),(30) as τi = 2παir a2c 3m . 8 (44) The scaling parameters are connected by the following relations t0, a0, m0, M0, ρ0, U0, Ω0, E0 t2 0 = a3 0 Gm0 , M 2 0 = Ga0m3 0, U0 = Gm2 0 a0 , ρ0 = m0 a3 0 , Ω0 = M0 m0a2 0 , E0 = U0a2 0, (45) so that in non-dimensional variables m = 4πa2cρ/3, M = 2ma2Ω/5. The total nondimensional energy Utot = Utot/U0 is written as ("tilde" is omitted) Utot = m(2 a2 + c2) 10 3m2 5a − ak + 5 4 M 2 ma2 + E (a2c)2/3 (46) 6 Equilibrium configuration and linear oscilla- tions In equilibrium there is no dissipation and the following relations follow from (35),(36), connecting the equilibrium parameters 25 4 M 2 m2a3 + 3 2 m a2 a1k + 10 3am E (a2c)2/3 = 0, m a2 a2k + 10 3cm − 3 E (a2c)2/3 = 0. (47) (48) At given m, M and E these equations determine the configuration of the sphe- roid. Finding a from (48) a = 10 9m2 E k5/3 a2k , (49) we obtain after its substitution into (47) the relation for determining k in the form E Taking into account (40), we come to the relation a1k + 2k a2k + 15 4 M 2 m k5/3 a2k = 0. k(1 + 2k2) (1 − k2)3/2 arccos k − 3k2 1 − k2 = 75 4 M 2 mE k5/3 a2k = 25 6 M 2 m3 k a . (50) (51) The last relation in (51) follows from the connection (49) between a and E, and in this form (51) coinsids exactly with the equation (9), determining the form of the Maclaurin spheroid with 9 a =(cid:18) 3m 4πkρ(cid:19)1/3 , of a given mass, density and angular momentum. In our case the density is determined by the entropy E. For the spheroid with unit non-dimensional mass m and semiaxis a we obtain in equilibrium relations determining M and E as functions of k 9 10 E = k5/3 a2k, M 2 = − 6 25 (a1k + 2k a2k). (52) The spheriod, described by ordinary differential equations for semiaxies a and c has two levels of freedom in oscilations corresponding to the first p- mode, and the first fundamental f -mode (see Cox, 1980). Consider (without loosing a generality) oscillations of a spheroid with m = a = 1. The E, Utot, m and M losses are quadratic to perturbations, so these values are conserved during linear oscillations. For the perturbations of δa and δc we have the following equations from (35),(36) δa = − 75 4 M 2δa + 3 2 δ(a1k) − 3 a1k δa − 10 E k2/3 δa − 20 9 E k5/3 δk − δa τrel , δc = −3δ(a2k) + 6 a2k δa − 10 E k5/3 δa − 50 9 E k8/3 δk − δc τrel . The perturbations are expressed via perturbations δa, δc as (53) (54) δ(ak) = −a2k δk, δ(a1k) = 1 + a2k + 2k a1k 1 − k2 δk, (55) δ(a2k) = 3k a2k − ak δk, 1 − k2 δk = δc − δa; τrel = 2παrelr k 3 . The equations (53),(54) with account of (55) may be solved analitically at weak damping for any k. Consider here for simplicity two limiting cases k = 1 (sphere), and k ≪ 1 (pancake). The sphere is related to the nonrotating body with M = 0. In this case we have also ak = 1, a1k = − 2 3 , a2k = 1 3 , E = 3 10 , δ(ak) = − δk 3 , δ(a1k) = 2δk 5 , δ(a2k) = − 4δk 15 . With account of (56) we have from (53),(54) the following equations δa = − 1 15 δk − δa − δa τrel , δc = − 13 15 δk − δa − δc τrel . (56) (57) These equations are completed by relation δk = δc − δa. Taking δa, δc ∼ exp(−iωt) we come to the characteristic equation 10 ω4 − 9 5 ω2 + 4 5 + iω τrel (cid:18)2ω2 − 27 15(cid:19) − ω2 τ 2 rel = 0 (58) The solution of a characteristic equation may be obtained analytically for a weak damping at ω ≫ 1/τrel. The 4 roots of this equation are ω1,2 = ±1 − i 2τrel , ω3,4 = ±r 4 5 − i 2τrel (59) The first root corresponds to a pure radial oscillations, and the second one describes the fundamental mode which does not change the volume and which is preserved in the non-compressible fluid. In fact, for a pure radial oscillations δa with δk = 0 we have δc = δa, and the equations (57) reduce to δa = −δa − τrel with a characteristic equation ω2 + iω is ω1,2 = − i to the first two roots in (59). The damping increments are equal for these two modes, what may be seen also from the numerical solution of (35)-(39) for a general nonstationary problem at large t (see next section). Note, that in the radial oscillation mode δa = δc, and in the fundamental mode δa ≈ −0.5δc. τrel − 1 = 0. The solution of this equation , which for a small damping obviously correspond 2τrel ±q1 − 1 4τ 2 rel For the case of a pancake k ≪ 1 we have from (43),(52), (55) 3π k5/3, M 2 = 25 , a2k = 1, E = , a1k = − 9 10 ak = π 2 π 2 , (60) δ(ak) = −δk, δ(a1k) = 2δk, δ(a2k) = − πδk 2 . With account of (60) we have from (53),(54) the following equations δa = − 3π 4 δa − 3δk − δa τrel , δc = − 5 k δk − 3δa − δc τrel , (61) with δk = δc − kδa. Taking δa, δc ∼ exp(−iωt) we come to the characteristic equation ω4 −(cid:18) 5 k + 3π 4 (cid:19) ω2 + 15π 4 + 6 + iω τrel (cid:18)2ω2 − 5 k − 3π 4 (cid:19) − ω2 τ 2 rel = 0 (62) The solution of a characteristic equation may be obtained analytically for a weak damping at ω ≫ 1/τrel. The 4 roots of this equation are i , ω3,4 = ±r 3π 4 − ω1,2 = ±r 5 k − In the pancake the oscillation modes (1,2) and (3,4) are almost independent on each other. In the first two modes δa ∼ kδc ≪ δc, what corresponds to rapid oscillations of the thickness of the pancake at almost fixed radius. Two other modes with a low frequency are characterized by relation δc ∼ kδa ≪ δa, 2τrel 2τrel (63) i 11 and correspond to the oscillation of the equatorial radius of the pancake at approximately the same thickness. Here both modes have the same damping increments because we have chosen the same relaxation time for the damping of two degrees of freedom, which may differ in reality for the flat pancake. The model, developed here may be easily generalized onto the 3 dimensional ellipsoids, where instead of analytical formulae for the gravitational potential and forces we'll have elliptical integrals. This model will have three degrees of freedom, with the additional one, corresponding to non-axisymmetric per- turbations. Contrary to the axisymmetric modes, which are always stable, the non-axisymmetric mode of the spheroid will become unstable at sufficiently large M , first due to a secular instability appearing only at presence of damping, and at larger M the spheriod will become dynamically unstable independently on the presence of the damping. In that case we may expect that in presence of damping the dynamical instability will appear at larger M , than in the pure conservative case (see Chandrasekhar, 1969). We suppose to investigate the three-dimensional model and its stability in the near future. 7 Numerical results The development of gravitational instability starts from almost Hubble expan- sion velocity which is transformed into contraction as a result of a developement of gravitational instability. To simplify the problem we start from a nonlinear phase of the instability with an expansion velocity smaller than the Hubble value. We start the simulation from the spherical body of the unit mass, small or zero entropy Ein ≪ 1, and negative total energy Utot < 0. We also specify the parameters, characterizing different dissipations αrel, αel, αml, αMl. The following parameters have been used in all variants of calculations ain = 1, cin = 1, min = 1, αrel = 3, αel = αml = αMl = 15. (64) Other four initial parameters ain, cin, M and E used in calculations are listed in the Table 1, where also the initial Utot is represented, which is calculated from (26). We have used in the calculations the Runge-Kutta code from (Press et al., 1992). The results of the calculations are very transparent, and are represented in figures 1-16. The odd number figures represent time dependences of a and c axies on time, together with the phase = sin(ph) = sin(R t 0 Ωdt), permitting the comparison between the oscillational and rotational motion. In the even number figures the entropy E, current mass m and current total energy Utot are given as functions of the time for the choosen set of the relaxation parameters (64). Here the entropy behaviour is the most interesting, because it is based on the realistic estimation of τrel. The characteristic times of other losses are more difficult to estimate, so they had been choosen equal to 5τrel, making these losses practically unimportant. After the initial expansion the collapse happens, during which the linear size decreases 10-40 times, and in the process of the subsequent relaxation the 12 N 1 2 3 4 5 6 7 8 ain 0.5 0.5 0.2 0.5 0.9 0.9 0.99 1.3 cin M 0.3 0.5 0.5 0.5 0.2 0.5 0.1 0.5 0.1 0.9 0.7 0.1 0.01 0.94 0.01 1.1 E 0.01 0.01 0.01 0 0 0 0 0 Utot -0.4025 -0.2025 -0.2655 -0.5125 -0.3445 -0.3765 -0.3155 -0.1409 Table 1: Initial conditions and parameters taken in numerical solution of equa- tions (35)-(39), in addition to the fixed values from (64); and initial value of the total energy, according to (26). object approaches the equilibrium, described by relations (52). The relaxation and entropy production, mass and energy losses happens in the most rapid stages at maximal compression, what is reflected in the steps of the functions E(t), m(t), Utot(t) , well visible in figures 8,10,12,14,16. The following remark- able properties should be noted. 1. For a low initial angular momentum, which is probably characteristic for the collapse of the primodial dark matter clouds, the collapse leads initially to the "pancake", in accordance of Zeldovich (1970) theory. On the subsequent stage of the expansion the spheroid becomes prolate, and during subsequent phases the spheroid is oscillating on the fundamental mode, changing the form between the prolate and oblate ones. It may be seen especially well in figures 11 and 15, where temporally the prolate figure is formed with c/a ∼ 2 − 5. We may expect the formation of such transient prolate figures in the central parts of the collapsing dark matter cloud, where the density may be taken as almost uniform, and where the characteristic period of oscillations is much less than the time of the whole body collapse. 2. The phenomenological relaxation time accepted in the calculations seems to be rather realistic (Lynden-Bell, 1967), so the calculated process of the damp- ing should be close to a reality. We have obtained, that oscillations preserve during more than 10 initial characteristic (Jeans) times tJ . That means, that in the most extended dark matter objects with the largest tJ the damping of oscillations is still in progress, and the dark matter objects may be presently in a nonstationary state. The collapsing dark matter ofjects and their oscillations could influence the visible picture of the fluctuations of the cosmic background microwave radiation, which are now under intensive investigation (see Naselsky, Novikov, Novikov, 2002). 3. The description of the dynamics of the spheroid by the system of equa- tions with two degrees of freedom give a possibility of the exitation of only two oscillation modes. In the case of a low rotation when two degrees of freedom 13 are strongly coupled the two modes reduce to the first radial p-mode, like the main mode of the radial oscillations of the sphere, and the fundamental f -mode, which almost does not change the total volume, and which survives in the non- compressible sphere (Cox, 1980). The lengths of two axies oscillate with the same period. Excitation of both modes may be distinctly seen in calculations. For some initial conditions the main remaining mode is the p-mode (figures 7,15); for others (figures 9,11,13) the fundamental f -mode was mainly excited. For rapidly rotating bodies, with a large equilibrium ratio of a/c, the oscilla- tions in two axies are almost independent, and their lengths are oscillating with substantially different periods (figures 1,3,5). 8 Emission of very long graviattional waves Gravitational radiation is produced during the collapse of the nonspherical body. Gravitational radiation during a formation of a pancake was estimated by Thuan and Ostriker (1974), and was improved by Novikov (1975), who took into ac- count the most important stage of the radiation during a bounce. The formula for the estimation of the total energy emitted during the collapse and bounce is UGW ≈ 0.01(cid:18) rg aeq(cid:19)7/2(cid:18) aeq cmin(cid:19) M c2. (65) Here aeq is the equilibrium equatorial radius of the pancake, and cmin is its minimal half-thickness during the bounce, rg = 2GM/c2 is the Schwarzschild gravitational radius of the body. From our calculations we have aeq/cmin ≤ 100. The value of rg/aeq we estimate using the observed average velocity of a galaxy in the rich cluster vg ∼ 3000 km/s, and taking rg aeq ∼(cid:16) vg c (cid:17)2 ≈ 10−4. (66) Than the energy carried away by the gravitational wave (GW) may be estimated as UGW ≈ 10−14M c2. If all dark matter had passed through the stage of a pancake formation, than very long GW with a wavelength of the order of the size of the galactic cluster have an average energy density in the universe εGW ≈ 10−14ρdmc2 ≈ 3 · 10−23erg/cm3. (67) Here we have used for estimation the average dark matter density ρdm = 3·10−30 g/cm3. The average strength EGW of the very long GW may be estimated, taking the relation (Landau and Lifshits, 1993) εGW = c2 16πG h2, (68) where h is metric tensor perturbation (nondimensional), connected with GW, we consider only the scalar, having in mind the averaged value of this perturbation. Taking into account h = ωh = 2πch/λ, λ ∼ 10 Mpc is the wavelength of GW of 14 the order of the size of the cluster. From the comparison of (67),(68) we obtain the averaged metric perturbation due to very long GW in the form 2λ ¯h = c2 (cid:18) GεGW π (cid:19)1/2 ≈ 6 · 10−11 (69) for the values of λ and εGW , mentioned above. Insidently we may expect 10 times larger amplitude of the GW, than the averaged over the volume value. We may estimate the angle of the lensing using its linear dependence on the potential difference, and the value of deviation 1.75 arc sec. for the Sun (Zeldovich and Novikov, 1971) which gravitational potential is equal to hsun = GM/Rc2 ≈ 2 · 10−6. Finally we get for the expected lensing influence of the very long GW θGW = θsun 10¯h hsun = 3 · 10−4θsun ≈ 5 · 10−4arcsec. (70) Such angular resolution is available by the very long base interferometry and the interferometric optical telescopes under construction could give much better angular resolution. 9 Conclusions The simplified model of the collapse and subsequent relaxation of the rotat- ing compressed shperoid may be related to the processes in the central parts ob the dark matter objects. The formation of the transient prolate dark mat- ter spheroid follows from the calculations for slowly rotating (or non-rotating) objects. For realistic values of the relaxation time we have obtained rather slow damp- ing of the oscillations, indicating that after ∼ 10 oscillations the amplitude may remain on the level of 1/10 of the Hubble value of the velosity for a correspond- ing mass scale. Such oscillations in the dark matter objects of the largest scale may be preserved until the present time. Interaction of the cosmic microvawe background radiation (CMB) with the dark matter objects on the stage of their collapse and subsequent oscillation phases may influence the visible characteristics of the observed picture of the fluctuations of CMB. The weak but very long gravitationl waves (GW) emitted mainly on the stages of the collapse and pancake formation form a long wave GW background, which also influence CMB fluctuations (Doroshkevich et al, 1977). Besides, the existence of such a long GW in the space between the source and the observer may be registered as an action of the gravitational lense. Absence of the appropriate lensing objects in the direction of any observed gravitational lense object may indicate to the existence of dark matter objects without the barion matter presence, or may be the indication to the presence of the very long GW which is moving in the space between us and the lensed object The expected angles of such lensing (∼ 10−3) arc sec. may be available in the near future. 15 1.5 1 0.5 1− 2− 3− a−axis c−axis phase 3 1 2 0 0 10 20 30 40 50 Figure 1: Time dependence of semiaxies a(t) and c(t), and sinus of the rotational phase: phase(t) for initial values of the variant 1, Table 1. The present approach describing the compressed spheroid by the two degrees of freedom may be evidently generalized for a description of the 3-axis spheroid by the ordinary equations with 3 degrees of freedom. In particular, it would be possible to investigate the influence of the damping processes on the bound- aries of the stability of the Maclaurin spheroid for the transformation into the Jacobi 3-axis ellipsoid (secular and dynamical) at large angular momentums. The model could be generalized also for the case of anizotropic relaxation and creation of anizotropic pressure. Acknowledgement I grateful to Theoretical Astrophysics Center (TAC), Copenhagen, for hos- pitality during the work on this paper, and to A.G. Doroshkevich and P.D. Naselsky for useful discussions. References [1] Chandrasekhar S. 1969, Ellipsoidal figures of equilibrium. Yale University Press. [2] Cox J.P. 1980, Stellar pulsation theory. Princeton Univ. Press. 16 1− 2− 3− entropy mass energy 1 0.5 0 2 1 3 −0.5 0 10 20 30 40 50 Figure 2: Time dependence of the non-dimensional entropy E(t), mass m(t), and total energy Utot(t) for initial values of the variant 1, Table 1. 3 2 e s a h p i , s x a − c , s x a − a i 1 0 0 1 2 − − 3 − a−axis c−axis phase 3 1 2 50 100 150 Time Figure 3: Same as in Fig.1, for initial values of the variant 2, Table 1. 17 1 0.5 0 y g r e n e , s s a m , y p o r t n e 1− 2− 3− entropy mass energy 2 1 3 −0.5 0 50 100 150 Time Figure 4: Same as in Fig.2, for initial values of the variant 2, Table 1. 2 1.5 1 0.5 1− 2− 3− a−axis c−axis phase 3 1 2 e s a h p i , s x a − c , s x a − a i 0 0 20 40 Time 60 80 Figure 5: Same as in Fig.1, for initial values of the variant 3, Table 1. 18 1 0.5 0 y g r e n e , s s a m , y p o r t n e 1− 2− 3− entropy mass energy 2 1 3 −0.5 0 20 40 Time 60 80 Figure 6: Same as in Fig.2, for initial values of the variant 3, Table 1. 1.5 1 0.5 0 0 1− 2− 3− a−axis c−axis phase 3 1 2 10 20 30 Figure 7: Same as in Fig.1, for initial values of the variant 4, Table 1. 19 y g r e n e , s s a m , y p o r t n e 1 0.5 0 −0.5 −1 0 1− 2− 3− entropy mass energy 2 1 3 10 20 30 Time Figure 8: Same as in Fig.2, for initial values of the variant 4, Table 1. 2 1.5 1 0.5 1− 2− 3− a−axis c−axis phase 3 1 2 e s a h p i , s x a − c , s x a − a i 0 0 20 40 Time 60 80 Figure 9: Same as in Fig.1, for initial values of the variant 5, Table 1. 20 1 0.5 0 y g r e n e , s s a m , y p o r t n e 2 1− 2− 3− entropy mass energy 1 3 −0.5 0 20 40 Time 60 80 Figure 10: Same as in Fig.2, for initial values of the variant 5, Table 1. e s a h p i , s x a − c , s x a − a i 2 1.5 1 0.5 0 0 1− 2− 3− a−axis c−axis c−axis phase 3 1 2 10 20 30 40 50 Time Figure 11: Same as in Fig.1, for initial values of the variant 6, Table 1. 21 1 0.5 0 y g r e n e , s s a m , y p o r t n e 1− 2− 3− entropy mass energy −0.5 0 10 20 30 Time 2 1 3 40 50 Figure 12: Same as in Fig.2, for initial values of the variant 6, Table 1. 2 2 1 1.5 1 0.5 1− 2− 3− a−axis c−axis phase 3 2 1 e s a h p i , s x a − c , s x a − a i 0 0 20 40 Time 60 80 Figure 13: Same as in Fig.1, for initial values of the variant 7, Table 1. 22 1− 2− 3− entropy mass energy 1 0.5 0 y g r e n e , s s a m , y p o r t n e −0.5 0 20 40 Time 2 1 3 60 80 Figure 14: Same as in Fig.2, for initial values of the variant 7, Table 1. [3] Doroshkevich A.G., Novikov I.D., Polnarev A.G. 1977, Astron. Zh. 54, 932 (Sov. Astron, 21, 932, 1977). [4] Doroshkevich A.G., Kotok E.V., Polyudov A.N., Shandarin S.F., Sigov Yu.S., Novikov I.D. 1980, Month. Not. R.A.S. 192, 321. [5] Doroshkevich A.G., Muller V., Retzlaff J., Turchaninov V. 1999, Month. Not. R.A.S. 306, 575. [6] Landau L.D., Lifshitz E.M. 1993, The classical Theory of Fields. Pergamon Press. [7] Lynden-Bell D. 1967, Month. Not. R.A.S. 136, 101. [8] Naselsky P.D., Novikov D.I., Novikov I.D. 2002, Relict radiation in the uni- verse. Nauka, Moscow (in Russian). [9] Novikov I.D. 1975, Astron. Zh. 52, 657 (Sov. Astron, 19, 398, 1976). [10] Press, W. H., Teukolsky, S. A., Vetterling, W. T., Flannery, B. P. 1992. Numerical recipes in FORTRAN. The art of scientific computing. Cambridge: University Press, 2nd ed. [11] Thuan T.X., Ostriker J.P. 1974, ApJ 191, L105. [12] Zeldovich Ya.B. 1970, Astrofizika 6, 119 23 5 4 1 e s a h p 3 2 1 1− 2− 3− a−axis c−axis phase i , s x a − c , s x a − a i 2 1 0 0 2 2 1 1 2 3 100 200 300 Time Figure 15: Same as in Fig.1, for initial values of the variant 8, Table 1. 1 0.5 0 y g r e n e , s s a m , y p o r t n e 2 1− 2− 3− entropy mass energy 1 3 −0.5 0 100 200 300 Time Figure 16: Same as in Fig.2, for initial values of the variant 8, Table 1. 24 [13] Zeldovich Ya.B., Novikov I.D. 1971, Relativistic Astrophysics, 1, Stars and Relativity, Univ. Chicago Press, Chicago. 25
astro-ph/9908077
3
9908
2000-01-05T23:51:25
Markarian 421's Unusual Satellite Galaxy
[ "astro-ph" ]
We present Hubble Space Telescope (HST) imagery and photometry of the active galaxy Markarian 421 and its companion galaxy 14 arcsec to the ENE. The HST images indicate that the companion is a morphological spiral rather than elliptical as previous ground--based imaging has concluded. The companion has a bright, compact nucleus, appearing unresolved in the HST images. This is suggestive of Seyfert activity, or possibly a highly luminous compact star cluster. We also report the results of high dynamic range long-slit spectroscopy with the slit placed to extend across both galaxies and nuclei. We detect no emission lines in the companion nucleus, though there is evidence for recent star formation. Velocities derived from a number of absorption lines visible in both galaxies indicate that the two systems are probably tidally bound and thus in close physical proximity. Using the measured relative velocities, we derive a lower limit on the MKN 421 mass within the companion orbit (R \sim 10 kpc) of 5.9 \times 10^{11} solar masses, and a mass-to-light ratio of >= 17. Our spectroscopy also shows for the first time the presence of H\alpha and [NII] emission lines from the nucleus of MKN 421, providing another example of the appearance of new emission features in the previously featureless spectrum of a classical BL Lac object. We see both broad and narrow line emission, with a velocity dispersion of several thousand km s^{-1} evident in the broad lines.
astro-ph
astro-ph
Markarian 421's Unusual Satellite Galaxy Peter W. Gorham 1, Liese van Zee 2,3 Stephen C. Unwin 1, and Christopher S. Jacobs 1 1. Jet Propulsion Laboratory, Pasadena, California, 91109 2. National Radio Astronomical Observatory, PO Box 0, Socorro, New Mexico ABSTRACT We present Hubble Space Telescope (HST) imagery and photometry of the active galaxy Markarian 421 and its companion galaxy 14 arcsec to the ENE. The HST images indicate that the companion is a morphological spiral rather than elliptical as previous ground -- based imaging has concluded. The companion has a bright, compact nucleus, appearing unresolved in the HST images. This is suggestive of Seyfert activity, or possibly a highly luminous compact star cluster. We also report the results of high dynamic range long-slit spectroscopy with the slit placed to extend across both galaxies and nuclei. We detect no emission lines in the companion nucleus, though there is evidence for recent star formation. Velocities derived from a number of absorption lines visible in both galaxies indicate that the two systems are probably tidally bound and thus in close physical proximity. Using the measured relative velocities, we derive a lower limit on the MKN 421 mass within the companion orbit (R ∼ 10 kpc) of 5.9 × 1011 solar masses, and a mass -- to -- light ratio of ≥ 17. Our spectroscopy also shows for the first time the presence of Hα and [NII] emission lines from the nucleus of MKN 421, providing another example of the appearance of new emission features in the previously featureless spectrum of a classical BL Lac object. We see both broad and narrow line emission, with a velocity dispersion of several thousand km s−1 evident in the broad lines. Subject headings: galaxies: BL lacertae objects: individual (Markarian 421) -- galaxies: binary -- galaxies: interactions 3presently at: Dominion Astrophysical Observatory, 5071 West Saanich Road, Victoria, B.C., V8X 4M6, Canada -- 2 -- 1. Introduction Markarian 421 is a giant elliptical galaxy that contains one of the nearest BL Lac objects, at a redshift of 0.0308 (Ulrich et al. 1975). This object is among the most intensively studied of all active galactic nuclei (AGN). MKN 421 is a strong cm-wavelength radio source that has shown reported superluminal motion in its compact radio jet (Zhang and Baath 1990), although recent measurements (Piner et al. 1999) do not confirm this. It is optically highly variable in both intensity and polarization. It is seen in X-rays (Comastri et al. 1997), GeV gamma-rays (Mukherjee et al. 1997), and up to multi-TeV energies (Krennrich et al. 1997). Episodes of rapid variability have been seen repeatedly at many wavelengths (Tosti et al. 1998; Gaidos et al. 1996) strengthening the evidence for the presence of a compact object as the source of the nuclear activity. The host galaxy of MKN 421 has been the subject of several spectroscopic and photometric studies. The first such study (Ulrich 1975) established the redshift based on weak stellar absorption lines, and also noted that a nearby galaxy 14 arcsec to the ENE had a similar redshift (z=0.0316), indicating that it was probably physically related, although if the velocity difference were due to the Hubble flow, the distance could be a few Mpc or more. The companion galaxy was classified as a normal elliptical (Hickson et al. 1984). Further work by Ulrich (1978) showed that MKN 421 was the brightest member of a group of 5 -- 7 galaxies of similar redshift spread over a region of sky of order 10 arcmin in radius. The presence of this group increases the likelihood that the companion's proximity to MKN 421 is physical rather than a random alignment. There is mounting evidence that AGN phenomena appear to be associated with galaxy mergers or encounters (cf. Shlosman, Begelman, and Frank 1990; Hernquist & Mihos 1995). In the case of BL Lac objects, a significant number have been found in the last decade or so to be associated with close companions or groups of nearby galaxies (cf. Falomo 1996; Heidt 1999), although MKN 421 has apparently been overlooked in this regard. Intrigued by the proximity of these two galaxies, we have analyzed archival HST imagery of the system, and performed Hale 5 m long-slit spectroscopic observations aimed at clarifying this association. Our goal was to understand the nature of the nearby galaxy in relation to MKN 421, and to investigate the properties of the companion galaxy itself. If the galaxy is as close to MKN 421 as its projected distance (∼ 10 Kpc) suggests, it is deep within the gravitational potential well of MKN 421, and is probably sweeping through its stellar halo. The conditions under which such an encounter can take place are of general interest in the understanding of galaxy evolution. Our results will show that the companion galaxy contains a Seyfert-like nucleus, and is likely to be tidally interacting with MKN 421. Although the evidence is circumstantial, this association does appear to lend weight to the -- 3 -- suggestions that galaxy encounters are an important factor in AGN evolution, and that close companions are associated in some way with the BL Lac phenomenon. In the following section we summarize the the imagery and photometry which indicate nuclear activity in the companion galaxy. Section 3 summarizes the spectroscopic results, from which we derive a velocity profile which indicates a system that is tidally bound. Section 4 presents some further analysis of the spectroscopic results, including a derived lower limit of the bulge mass and mass -- to -- light ratio of MKN 421 under some plausible assumptions. Section 5 summarizes and concludes the article. 2. Observations A complete log of all observations presented here is shown in Table 1. Table 1: Log of observations presented here. Date Telescope Instrument Filter/grating 1997 March 5 1999 February 19 Palomar 5m Double Spect. 1999 May 24 1999 May 24 WF/PC2 WF/PC2 WF/PC2 HST HST HST F702W 600/300 l/mm F555W F814W In the work of Ulrich (1978), the system 14′′ to the ENE of MKN 421 was denoted as galaxy number 5 of the group of galaxies associated with MKN 421. Although the galaxy had been noticed prior to this work, it appears that Ulrich was the first to provide a designation for it. Thus we will refer to it in this present work as MKN 421-5. We note that users of NASA's Extragalactic Database (NED) may find a reference to this object as RX J1104.4+3812:[BEV98] 014. However, since this reference is only used to distinguish objects within the ROSAT error circle for MKN 421, and the object is previously known, we prefer the designation given by Ulrich (1978). 2.1. HST Imagery and Photometry -- 4 -- 2.1.1. HST images MKN 421 was observed with the HST wide field/planetary camera (WF/PC2) on 1997 March 5, using the F702W filter. Five exposures, one of duration 2 s, two of duration 30 s, and two of duration 120 s were made 2. The images were prepared by standard techniques described in Holtzman et al. (1995), and the moderate cosmic ray contamination of the frames was repaired by hand using linear interpolation. In the 2 s exposure, the MKN 421 nucleus is not saturated, but the companion galaxy is underexposed. The remaining two images overexpose the AGN, but for both MKN 421's host galaxy and the companion galaxy the exposures provide good signal- to-noise ratios over the sky background. The pixel scale in these images is 0.0455 arcsec per pixel, and the resolution of HST at the mean filter wavelength (690 nm) is ∼ 0.080 arcsec; thus the images are sampled just slightly under the Nyquist frequency. Secondary corrections, including those associated with the known gradient in charge transfer efficiency and the pixel area variation across the frame (Holtzman et al. 1995), are not corrected for in the displayed images, but we have applied corrections for these effects in the relative photometry presented in the next section. MKN 421 was again observed with WF/PC2 on 1999 May 24 with both the F555W and F814W filters in snapshot mode, with 300s exposures in each case.3 Since these were single frame integrations, the cosmic ray removal for presentation purposes is more difficult, and the images do not add significantly to the structure seen in the F702W image. Thus we do not present these as part of the imagery here, although we do include results of the photometry performed on these additional images in the following section. Figure 1 shows a slightly smoothed grayscale of the summed F702W image, with a logarithmic stretch and quantized levels chosen to show the details of the host galaxy of MKN 421 and MKN 421-5. Several features of MKN 421-5 are evident even from this image. First, its structure is not a simple elliptical. A suggestion of spiral arms is evident, and possible evidence of barlike structure appears at the outer edges of the galaxy. Second, the nucleus of the companion is clearly brightened relative to the galactic bulge, as is shown in the inset frame. Third, there is no evidence for any obvious dust lanes or similar absorption features in either the outer or bulge regions of the galaxy. 2The HST observations used here are available as part of the Space Telescope Science Institute public archive, and were made originally as a result of a proposal by C. M. Urry. 3These HST observations are also available as part of the Space Telescope Science Institute public archive, and were made originally as a result of a proposal by R. Fanti. -- 5 -- Fig. 1. -- HST image of MKN 421 and companion MKN 421-5, through red F702W filter, 300 s total exposure. In the larger image the cores of MKN 421 and the companion are saturated to show detail of the companion disk structure. Levels are quantized to more clearly show the companion structure. The inset shows the nucleus of the companion plotted with a logarithmic stretch, showing the bright, unresolved nucleus. The scale is approximately 700 pc/arcsec for h = 0.65. -- 6 -- Fig. 2. -- (a) Logarithmic grayscale of a smoothed HST image of MKN 421-5. Central portions saturated to show disk more clearly. (b) Plot of surface brightness centroids of slices perpendicular to the major axis showing evidence of spiral structure. -- 7 -- Figures 2 -- 3 illustrate these features. In Figure 2 we show more quantitative evidence for the presence of spiral structure, since the structure is difficult to reproduce adequately in paper copies of the images themselves. Fig. 2(a) is a contour plot centered on the companion galaxy showing structure which is suggestive of spiral arms. In Fig. 2(b) we show centroids of the surface brightness distributions in slices along the galaxy major axis (dashed line) in Fig. 2(a). The displacement of the brightness centroids clearly shows the presence of spiral structure. (The amplitude of the centroid displacement does not directly track the location of the arms, since it depends on the brightness of the arm structure relative to the surrounding disk.) In Figure 3 we show a profile of relative surface photometry using the F702W filter along the major axis of the companion galaxy. We have averaged the surface brightness in opposite ≃ 15◦ sectors, and have performed one-dimensional fits to 3 components that are apparent in the image: an unresolved core; a bulge component, here represented by a Gaussian; and an exponential disk. We have not truncated the disk at an inner radius here; our goal is primarily to show that the data are not consistent with a single power-law or other one-parameter model for radial brightness, but are reasonably well-modeled by the standard parameters of a disk system. 2.1.2. HST photometry We estimated photometric parameters from the HST images using the prescription given by Holtzman et al. 1995, including all of the relevant corrections. We have also transformed the resulting magnitudes to Cousins VRI magnitudes using the transformations in Holtzman et al., and we include also the transformed colors although we note that the V and I measurements were not contemporaneous with the R measurements. In Table 2 we summarize the results for the nuclear magnitudes of the two galaxies, and the annular bulge magnitude of MKN 421-5. We have also converted the measured magnitudes to absolute magnitudes using h = 0.65, for use in later discussion. For MKN 421, we show only an estimate of the R magnitude of the nucleus for comparison with MKN 421-5; in the 1999 May observations in the F555W and F814W filters there were no exposures short enough to avoid saturation of the nucleus. Based on the expected contributions of the different components indicated by Fig. 3, The nuclear magnitude may contain a significant bulge contribution, of order 30-40%. In spite of this, it still appears that the bright nucleus is too compact to be a typical star forming region. As we discuss in a later section, it is likely to be either a compact nuclear star cluster or more probably an active nucleus. And although there is only a -- 8 -- Fig. 3. -- Relative surface photometry of 15◦ sectors centered on the MKN 421-5 major axis, as a function of radial distance from the galaxy center. We also show best-fit curves for the relative contributions of an unresolved nuclear source, a surrounding bulge (taken to be a Gaussian), and an exponential disk (with no inner truncation). The need for mutiple components in the fit strengthens the case that this is actually a disk system. -- 9 -- Table 2: HST photometry of MKN 421 & companion. Units are magnitudes, and standard errors are typically 0.04 mag. No unsaturated V or I band exposures of the nucleus of MKN 421 were available. Source MKN 421 AGN MKN 421-5 nucleus (r ≤ 0.18") MKN 421-5 bulge (0.18′′ ≤ r ≤ 1.1") V ... 19.66 19.62 R 12.74 19.11 19.05 I ... 18.53 18.41 V − R V − I MR ... 0.56 0.58 ... 1.14 1.21 -23.0 -16.6 -16.7 marginal difference in color between the nucleus and the annular bulge region, it does favor a rise of the nucleus relative to the bulge in the near -- infrared, which is consistent with Seyfert -- nucleus behavior. The nucleus of MKN 421 itself is highly variable in the visual bands. In fact, the HST observations in early 1997 occurred shortly after a large optical outburst (Tosti et al. 1998) during which the R-band magnitude peaked at brighter than 12. During March 1997, Tosti et al. estimated R = 12.4 from ground-based photometry. The HST observations, taken approximately 2 months later, show a decrease of 0.34 mag. 3. Optical Spectroscopy Low resolution optical spectra of MKN 421 and its companion were obtained with the Double Spectrograph on the 5m Palomar4 telescope during the night of 1999 February 19. The long slit (2′) with a 2′′ aperture was centered on the companion galaxy and two 600 sec exposures were obtained. The slit was positioned at an angle of 53◦, and passed through both the companion galaxy and the nucleus of MKN 421. A 5500 A dichroic was used to split the light to the two sides (blue and red), providing nearly complete spectral coverage from 3600 -- 7600 A. The blue spectra were acquired with the 600 lines/mm diffraction grating (blazed at 4000 A). The red spectra were acquired with the 316 lines/mm diffraction grating (blazed at 7500 A). Thinned 1024×1024 Tek CCDs, with read noises of 8.6 e− (blue) and 7.5 e− (red), were used on the two sides of the spectrograph. Both CCDs had a gain of 2. e−/(digital unit). The effective spectral resolution of the blue camera was 5.0 4These observations at the Palomar Observatory were made as part of a continuing cooperative agreement between Cornell University and the California Institute of Technology. -- 10 -- A (1.72 A/pix); the effective spectral resolution of the red camera was 7.9 A (2.47 A/pix). The spatial scale of the long slit was 0.62 ′′/pix for the blue camera and 0.48 ′′/pix for the red camera. The spectra were reduced and analyzed with the IRAF5 package. The spectral reduction included bias subtraction, scattered light corrections, and flat fielding with both twilight and dome flats. The 2 -- dimensional images were rectified based on arc lamp observations (Fe and Ar for the blue side; He, Ne, and Ar for the red side) and the trace of stars at different positions along the slit. The sky background was removed from the 2-dimensional images by fitting a low order polynomial along each row of the spectra. One dimensional spectra of MKN 421 and its companion were extracted from the rectified images using a 1.5′′ extraction region (the seeing disk at the time of the observations) centered on the peak emission of each system. In addition, a galaxy spectrum for MKN 421 was obtained by averaging together two 4′′ regions offset from the nucleus by ±4′′. While the night was non -- photometric, observations of standard stars from the list of Oke (1990) provided calibration for the fluxes which are estimated to be accurate to ∼ 10%. The flux -- calibrated spectra are presented in Figure 4, in units of erg cm−2 s−1 A−1. We estimate an overall uncertainty of 20% in the flux calibration due to variations in the transparency during our observations. 3.1. Spectrum of MKN 421 As seen in Figure 4a, the nucleus of MKN 421 is dominated by nonthermal emission with very weak absorption features. This new spectrum is similar to other observations of the nucleus of MKN 421 (e.g., Marcha et al. 1996), with the exception that [OI], [NII], and Hα emission lines have been detected. While measurement of relative fluxes for the narrow emission features is complicated by the presence of broad Hα emission (Figure 5), and by the contamination of Hα absorption from the underlying stellar population (see Figure 4b), the [NII] lines are significantly stronger than the narrow Hα feature. Large [NII]/Hα ratios are not uncommon in AGN, however (e.g., Veilleux & Osterbrock 1987). Both the broad and narrow emission lines appear to be associated only with the nucleus of MKN 421. In Fig. 5 we plot a fitted continuum model for MKN 421 along with the measured spectrum, in the region near the detected emission lines. Although the continuum model does not account for all of the structure in the spectrum near the emission lines, it does qualitatively indicate the instrinsic width of the broad-line emission, which has a 5IRAF is distributed by the National Optical Astronomy Observatories. -- 11 -- Fig. 4. -- Palomar 5 m double spectrograph spectrum of MKN 421 and its companion. The slit was aligned with the two nuclei. Units are erg cm−2 s−1 A−1 with an overall uncertainty of 20% in the flux calibration. (a) MKN 421 shows weak emission at Hα and NII, the first time any emission lines have been noted for this usually almost featureless BL Lac object. (b),(c) MKN 421 and its companion galaxy have a number of well-defined absorption lines but there is no detectable emission from either galaxy. -- 12 -- Fig. 5. -- Emission lines seen in MKN 421. The curve shows a power-law continuum model fitted to the overall spectrum. Both broad and narrow emission line components are evident above the continuum. A typical statistical error bar for the individual points is shown at the upper right. -- 13 -- full-width-at-half-maximum (FWHM) of order 80 A with detectable emission that extends out to nearly twice this value. The implied velocity dispersion is ∼ 5000 km s−1. We discuss some of the implications of this in a later section. Recently, Morganti et al. (1992) obtained narrow band images roughly centered on the Hα and [NII] lines and reported a total Hα+[NII] flux of 1.6 × 10−14 erg s−1 cm−2 for MKN 421. The present observations are the first spectroscopic detection of emission lines associated with the nucleus of MKN 421, confirming this result. The total integrated flux in the broad and narrow emission lines is 1.7 × 10−14 erg s−1 cm−2 (EW ∼ 2.2 A). We estimate an overall uncertainty in this value of ∼ 20%. This apparent agreement with Morganti et al. is somewhat misleading, since our estimate includes flux spread over more than 100 A, while the Morganti et al. estimate was based on a 50 A FWHM filter which was apparently offset with respect to the centroid of the emission. Thus we actually have detected a significantly lower total flux level than Morganti et al., although we cannot quantify the difference without a more accurate knowledge of the filters used by Morganti et al. Despite the new detection of emission lines, MKN 421 still falls under the general class of BL Lac objects: the 4000 A break has a low contrast in the nuclear regions and the derived equivalent widths for the emission lines are significantly less than 5 A. The sudden appearance of emission features in previously featureless spectra of BL Lac -- type objects is becoming quite common. Just a few years ago, such features were detected in the prototypical BL Lac, BL Lac itself, for the first time (Vermeulen et al. 1995; Corbett et al. 1996). Similar events were discovered in OJ 287 (Sitko & Junkkarinen 1985) and PKS 0521 -- 365 (Ulrich 1981; Scarpa, Falomo, & Pian 1995) as well. The rise and decline of emission -- line features in BL Lac -- type objects poses an interesting puzzle, and monitoring of these lines should be done as often as possible to determine if their intensity is correlated to that of other bands. 3.2. Spectrum of the Host Galaxy A spectrum of the underlying host galaxy was obtained by averaging spectra on either side of the nucleus (Figure 4b). As found in other spectroscopic observations of the host galaxy (e.g., Ulrich et al. 1975; Ulrich 1978), the spectrum is dominated by absorption features. In addition to the usual strong absorption features in the blue, the Na I λλ5890,5896 doublet is remarkably strong throughout the galaxy. We note that the presence of an apparent broad emission bump in the region of the spectrum surrounding -- 14 -- the Hα absorption line is apparently due to scattered emission from the extremely bright active nucleus of MKN 421. 3.3. Spectrum of MKN 421-5 Similar to the host galaxy of MKN 421, the spectrum of the companion galaxy is dominated by absorption lines (Figure 4c). However, the presence of strong Balmer absorption features indicates that this system is not dominated only by an old stellar population; rather, it must have had recent star formation activity within the last Gyr or so. The plethora of absorption lines in the optical spectrum permit an accurate redshift determination for this system: 9380 ± 50 km s−1. Ulrich (1978) was one of the first to point out that MKN 421 is part of a large group of galaxies; we now know that radio galaxies tend to form in groups (Zirbel 1997). The close proximity (spatially and in velocity space) of MKN 421 and MKN 421-5 suggests that these two systems may have had significant tidal interactions in the past and may explain the unusual stellar population of the companion. 3.4. Spatially resolved Na absorption Both MKN 421 and its companion galaxy have high signal -- to -- noise Na I absorption features which can be used to trace their gas kinematics. A position -- wavelength diagram for the two systems is shown in Figure 6. The Na line can be traced almost continuously from MKN 421 to MKN 421-5. The mean velocity as a function of position along the slit for several of the absorption lines is shown in Figure 7. The errors in each individual measurement are largely due to the relatively low signal -- to -- noise ratio and the somewhat coarse spectral resolution of the observations; nonetheless, the two systems are clearly offset in velocity, with a sense of velocity continuity between the two galaxies. Fig. 7 also displays the results of unconstrained least-squares fits for the velocity gradients across each of the two systems. The results of these fits are shown in Table 3. The velocities are not corrected for earth orbital motion; however, the correction for the hour angle of our observation is negligible compared to the errors in the estimates in Table 3. -- 15 -- Fig. 6. -- Spatially resolved absorption line profile for the NaI doublet. The centroid of the doublet shows a velocity trend that smoothly joins the two galaxies, an indication of a tidal interaction. -- 16 -- Fig. 7. -- (a) Contour plot of Palomar 60" Gunn r-band image of MKN 421 and companion. The heavy dashed line marks the slit position for the absorption line spectroscopy. (b) Plot of line centroids as a function of slit position. The lines show the fitted velocity gradients across each galaxy separately. The velocity gradient of MKN 421 is consistent with a component of Keplerian rotation that is in the same sense as the companion's velocity relative to MKN 421. The two thus appear to be a bound pair. -- 17 -- Table 3: Fitted velocities and velocity gradients for MKN 421 & companion. Source ¯v σv ∇v σ∇v km s−1 km s−1 km s−1 as−1 km s−1 as−1 MKN 421 MKN 421-5 8884 9380 27 50 23.0 7.4 6.0 4.0 3.5. On limits to radio emission from MKN 421-5 A number of VLA studies of MKN 421 and its immediate vicinity were conducted soon after it was recognized to be a BL Lac object (Ulvestad et al. 1984; 1983); however, the resolution of these studies was not adequate to show the companion due in part to the high dynamic range required (MKN 421 is 0.57 Jy at 20cm). More recent 20cm observations (Laurent-Muhleisen et al. 1993) with higher resolution show complex extended emission in the vicinity of MKN 421, but probably do not have enough dynamic range to detect the companion galaxy nucleus, if its brightness is comparable to the average of its class. Although there are conspicuous examples of radio -- loud Seyferts (primarily Seyfert 1) more typical Seyfert nuclei of both type 1 and 2 emit total fluxes of order 1021 W Hz−1 at cm wavelengths (cf. Ulvestad and Wilson 1989 and references therein). At the distance of the companion, this luminosity would produce about 0.5 mJy of 3.5 cm flux density, and a factor of 2-3 more than this at 20 cm for a typical Seyfert spectral index of −0.5 to −0.7. Inspection of a number of VLA maps at various resolutions shows no source of significant strength at the companion galaxy position. However, the complex halo of emission around MKN 421 precludes identifying a source at a level below 1 mJy. Further radio observations should be made to attempt to identify any compact source associated with the companion galaxy nucleus. 4. Discussion 4.1. The nature of the companion The absolute magnitude of the nucleus of MKN 421-5 was given in Table 2 as MR = −16.6. If we assume this magnitude contains a 40% background contribution from the galaxy of MKN 421-5, the nucleus still has MR = −16.0, and a corresponding -- 18 -- MV ≃ −15.5. It is thus apparently as bright or brighter than any of the compact nuclear clusters described from HST observations by Phillips et al. (1996) and Carollo et al. (1997) which fell mostly in the range of MV = −12 to −14. A recent detailed HST study of four nearby spirals with compact nuclei (Matthews et al. 1999) gave MB values that range from -8.5 to -10.4, and corresponding MI values of -9.9 to -11.4. Compact star cluster nuclei are not uncommon among late-type spiral galaxies, although they are often difficult to detect in ground-based images. However, based on the photometric results presented above, the nucleus of MKN 421-5 must be among the most luminous in its class if it comprises an overluminous nuclear cluster. The alternative is that we are viewing a Seyfert nucleus near the low end of the Seyfert luminosity range. In fact, recent studies have shown that the Seyfert luminosity function extends well below the luminosity of MKN 421-5 (Ho et al. 1996, Ho et al. 1997). It is also worth noting that Malkan et al. (1998), in an HST survey targeting known or potential Seyfert galaxies, found that the presence of a bright, unresolved nucleus in HST images was a nearly perfect indicator of Seyfert activity in a survey of several hundred objects. Thus the lack of emission lines from the nucleus of the companion galaxy is puzzling, given the other indications that the source is non-thermal in nature. However, given the mediocre seeing conditions during the spectroscopic observations, and the extreme compactness of the core, our constraints on emission lines are not yet very strong. The HST results indicate that we are viewing the companion at a substantial inclination relative to its spiral axis. This could contribute to the lack of observed emission lines due to obscuration. If the galaxy is a Seyfert 2, the narrow line emission may in fact be sufficiently diluted by the galactic bulge that it will be difficult to observe from the ground. From the HST image, the nuclear source is probably unresolved; even if we assume its diameter to be 0.1", the dilution factor under our seeing conditions (1.2") is at least a factor of 10, after accounting for the brightness of the nucleus relative to its surrounding bulge. Clearly, interpretation of any ground-based observations of this and other similar objects must be tempered with caution because of the danger of these dilution effects. 4.2. Implications of the velocity measurements 4.2.1. Mass of MKN 421 bulge As noted above, the shape of the velocity curve is consistent with the behavior expected from tidal interaction. This, combined with the proximity of MKN 421-5 to MKN 421, and the presence of a group of galaxies of similar redshift, leads to the conclusion that MKN -- 19 -- 421-5 is undergoing an orbital encounter with MKN 421. Also, since MKN 421-5 is much closer to MKN 421 than any other galaxy in the group, we may treat the pair as a binary system for purpose of estimating the dynamical masses involved. The statistics of binary galaxies have been treated in detail by Noerdlinger (1975) who showed that the assumption of circular orbits for a binary pair is a conservative one with regard to mass estimation. In fact, the circular orbit approximation is valid up to eccentricities of ǫ ∼ 0.4, and underestimates the mass by only a factor of 3 in the limit of a parabolic encounter (ǫ = 1). Thus we estimate a lower limit to MKN 421's mass M within the companion's present projected orbital radius Rp: M(Rp) ≥ U 2 r Rp G (1) where Ur is the companion's radial velocity and G is the gravitational constant. For Ur = 496 km s−1 and Rp = 10 Kpc in the comoving frame, the lower limit of the mass is M(Rp) ≥ 5.9 × 1011 M⊙. (2) For moderate orbital eccentricity (ǫ ≤ 0.4), Noerdlinger (1975) has shown that the most probable values of the total velocity UT and true separation RT are likely to be 20-30% higher than the observed values. Thus the most probable value for the bulge mass is about twice the value above, of order 1012 M⊙. Typical giant elliptical rotation curves increase out to radii of order 10 kpc, then flatten and continue flat out to 50 kpc or more. Since a flat rotation curve implies a linear increase of the dark halo mass with radius, we expect that only of order 10 -- 20% of the total mass of the MKN 421 system is likely to reside within Rp, the total mass of the galaxy probably approaches 1013 M⊙, among the highest masses of giant elliptical galaxies. Further measurements of the virial velocities of other members of MKN 421's group should be able to confirm this. 4.2.2. Mass -- to -- light ratio The CCD photometry of MKN 421 reported by Kikuchi et al. (1987) gives the absolute V magnitude of the MKN 421 host galaxy within a 13′′ radius of MV (13′′) = -- 21.51 ± 0.03, implying a luminosity of 3.5 × 1010 L⊙. Note that the emission from other bands, such as radio continuum and X -- ray, is almost entirely associated with the AGN rather than the galaxy. Thus, the implied minimum mass -- to -- light ratio is M/L ≥ 17, indicating a substantial dark matter contribution to the total mass of MKN 421 within Rp. -- 20 -- This calculation ignores the possible contribution of atomic or molecular gas to the total mass, but no HI has been detected in this system (van Gorkom et al. 1989). IRAS measurements of far-infrared fluxes of MKN 421 (van Gorkom et al.) have been interpreted as due to the presence of a total dust mass of order 5 × 107 M⊙. If a gas component were present corresponding to this dust level, one might expect up to 109 M⊙ of gas in the total galaxy; only a fraction of this would be within Rp. Thus we expect our derived M/L ratio to be robust. A number of statistical estimates of the mass -- to -- light ratios of binary galaxies have yielded values with ranges of 12 -- 32 for pure spiral pairs (Schweizer 1987; Honma 1999) and 22 -- 60 for elliptical pairs (Schweizer 1987). These estimates use galaxies with mean separations of order 60 -- 100 Kpc, and thus include a much greater mass fraction of the whole galaxy than our estimate. As discussed above, typical giant elliptical rotation curves lead one to expect that the total M/L for MKN 421 may be 5-10 times higher than what is measured for this close pair. 4.2.3. Estimates of MKN 421 central black hole mass Wandel (1999), Wandel et al. (1999), Laor (1998) and others have noted the correlation of AGN host galaxy bulge mass Mbulge to the mass of a central black hole MBH in cases where a reliable virial mass or reverberation mapping mass could be estimated. Wandel (1999) finds a relation of MBH ≥ 3 × 10−4 Mbulge in a sample composed primarily of low-luminosity AGN for which the statistics are reasonably good. Magorrian et al. (1998), in a comprehensive study of kinematics of normal galaxies, found a relation MBH ≥ 6 × 10−3 Mbulge. If we use these values to bound the likely value of MBH for MKN 421, and assuming that our estimate of the lower limit to the mass within Rp is dominated by the bulge mass of MKN 421, we expect a central black hole mass in the range 1.8 × 108 ≤ MBH ≤ 3.6 × 109 M⊙. We can also use the parametric relations given by Laor (1998) and Wandel et al. (1999) to make an independent estimate of the size of the broad-line region (BLR) based on the AGN bolometric luminosity. Wandel et al. give a black hole mass estimate based on the size of the BLR and ∆vF W HM , the FWHM of the measured broad-line velocity dispersion: MBH ≃ 1.45 × 105M⊙ cτBLR 1 light day!(cid:18) ∆vF W HM 1000 km s−1(cid:19) 2 . The size of the BLR can be estimated from (Kaspi et al. 1997; Wandel et al. 1999): rBLR = 15L 1/2 44 light days. (3) (4) -- 21 -- where L44 is the bolometric liminosity For MKN 421, estimates of the bolometric luminosity over the 0.1 to 1 µm band vary considerably, and it is unclear whether the quiescent or high-state luminosity is more appropriate. Using a bolometric luminosity derived from our HST magnitudes and the conversion given in Laor (1998), we find L44 ∼ 40, which gives rBLR = 90 light days. Assuming that Hα is dominant in forming the broad line emission, then we use ∆vF W HM ≃ 3300 km s−1 from our spectrum above. The implied black hole mass is then MBH ∼ 1.4 × 108 M⊙ which is of of the same order of magnitude as the lower limit estimate from the bulge mass above. A black hole mass of 109 M⊙ is of order 0.25% of the total dynamical mass out to 10 Kpc. Thus it is not in itself enough to affect the large scale dynamics between the two galaxies, although it would certainly strongly influence the inner bulge regions. It is notable, however, that the center of mass of the system will likely be significantly displaced from the position of the black hole, perhaps by a few hundred pc or more, depending on the companion mass and the dark matter distribution. Careful measurements of the virial motion of the stars near the center of MKN 421 may be able to detect such an offset, which would provide independent confirmation of the proximity of MKN 421-5 to MKN 421. 4.2.4. The dynamical state of the system The smooth rise in the rotation curve of MKN 421 out to Rp is consistent with expectations for a post-encounter system, in which any major velocity distortions produced by the encounter have relaxed and bulk tidal motions now dominate. However we can not rule out the possibility that this is the companion's first approach toward MKN 421. It is also possible that there are distortions in the velocity curves that are unresolved at our present sensitivity. Since we cannot unambiguously infer the companion's true velocity from radial velocity measurements alone, we cannot prove that the system is bound, and we can only base our estimates of the tidal effects on the global properties we observe. Despite these caveats, there are a number of different lines of evidence that point to a bound and tidally relaxed system. An obvious, though not compelling, piece of evidence is the fact that we observe the companion galaxy this close to MKN 421 at all, at the relatively late epoch of the system. The probability that we are observing a first encounter between two high-ranking cluster galaxies cannot be estimated based on this single case alone. However, based on the galaxy count of Ulrich (1978) and on our own Palomar images over a 12 arcmin field around MKN 421, there are potentially 7 to 10 galaxies of comparable brightness to MKN 421-5, in a volume of order 6 × 1016pc3, of order 104 times the volume containing MKN 421 and its -- 22 -- companion. Since MKN 421 will probably tend to significantly perturb the local potential, the chance of near encounters with it are not determined simply by chance; however, we estimate that the chance observation of such an encounter is of order 1% or less for any snapshot of similar clusters. If the system is not in the initial phase of a first encounter, it follows that (a) it is likely to be bound (due to the mass dominance of MKN 421 and the narrow range of phase space required for MKN 421-5 to have the required escape velocity); and (b) that it is tidally relaxed, since the relaxation time for the system as observed is in the range of 100 Myr, relatively short compared to the time of cluster formation. The evidence for recent star-formation in the companion galaxy is also relevant to discussion of the dynamics of this pair. Recent simulations of hierarchical galaxy encounters (cf. Bekki 1999 & references therein) have shown that star forming activity in the satellite galaxy is a likely result of such an interaction. A number of recent HST studies have found a significant number of companions near QSO host galaxies (Bahcall et al. 1995; Disney et al. 1995) which appear to be undergoing some type of interaction with the QSO host. Canalizo and Stockton (1997) found that at least one companion galaxy in such a system (PG 1700+518) showed evidence of a relatively young stellar population. What is perhaps surprising in the case of MKN 421 and MKN 421-5 is that, although simulations such as those of Bekki (1999) indicate that the companion galaxy is likely to evolve to a dwarf elliptical or irregular as a result of the encounter, MKN 421-5 appears to in fact be a relatively normal spiral, without any apparent major disruption or irregularity. This system thus presents a challenge to models which cannot account for preservation of such structure even in what was probably a strong tidal encounter between these two systems. It is notable that MKN 421-5, as a probable Seyfert, is likely to have a massive compact object in its nucleus. Based on arguments similar to the previous section and our estimates of the bulge luminosity of MKN 421-5, the expected mass is in the range of 2 − 5 × 106 M⊙. The presence of this large central mass may in fact have important dynamical consequences in stabilizing a satellite galaxy in such an encounter. 4.2.5. A schematic geometry for the system To postulate a possible three-dimensional geometry for this pair of galaxies, we combine a number of related pieces of evidence that bear on the geometry, including our measured velocity curves, the apparent inclination of the companion spiral, the major and minor axes of MKN 421, and parameters associated with the central mass of MKN 421, although the -- 23 -- latter are included mainly for comparison rather than as constraints on the global geometry. The position angle of the center of the companion relative to MKN 421 is 53◦, and we estimate that the projected major axis of the spiral lies at a position angle of 65◦. Thus the projected rotation axis (perpendicular to the major axis of the spiral) of MKN 421-5 is at a position angle of 335◦. The inclination of the spiral is more difficult to estimate, since the shape may be confused by an inner bar feature not easily distinguished in the images. However, it appears that the spiral inclination is in the range of 45◦ to 60◦ to the line-of-sight. If we assume that the disk of the spiral and the orbital plane are roughly coincident (this is reasonable if the systems are tidally relaxed), then we can make a first order schematic estimate of the overall geometry of the system, if the two galaxies are bound in an orbit of only moderate eccentricity. This schematic view of the system is depicted graphically in Figure 8. We have not attempted to quantify this geometry further; we present it simply as a qualitative baseline which can be refined (or revised) by future observations. Recent VLBI imagery and estimates of the jet doppler factors from both VLBI and gamma-ray observations (Piner et al. 1999; Gaidos et al. 1996) indicate conclusively that the jet angle must be ≤ 5◦ with respect to the observation vector. The projected position angle of the inner jet from VLBI observations is ∼ 320◦. We have plotted this position angle in Fig. 8 for comparison. For the case of a small angle with respect to the line-of-sight, the approximate coincidence of the projected position angle with that of the spiral axis and the possible orbital plane is presumably an accident. However, if the jet is actually physically aligned with the other angular momentum vectors in the system, then a much larger angle for it with respect to the line-of-sight is favored. An outstanding question in MKN 421 and in BL Lac objects in general is how well-aligned the jet axis is with the line-of-sight to the AGN. If the results based on estimates of the doppler factors are correct, and the jet axis is inclined by ≤ 5◦, then we are led to conclude that either the orbital plane of the two galaxies are not well-aligned with the perpendicular to the jet axis, or plane of the spiral companion is not well-aligned with the overall orbital plane. In either case the residual torques may have observable consequences, either in precession effects or tidal velocity distortions of the system. More complete velocity mapping may provide information on the latter consequence. As to the former effect, further analysis is necessary to estimate whether there might be observable effects, since the precession time scale would be expected to be long compared to the orbital time scale. -- 24 -- N MK 421-5 spiral axis far side spiral rotation E radio jet projected axis near side minor axis orbital motion Fig. 8. -- A suggested schematic geometry for the MKN 421 + MKN 421-5 system. MK 421 -- 25 -- 4.2.6. Companion rotation: Prograde or Retrograde? We conclude this section by briefly noting the sense of the companion rotation. The rotation curve shown in Fig. 7 shows that the eastern end of MKN 421-5 is redshifted relative to the western end which is closest to MKN 421. This indicates a rotation which is clockwise on the sky, and corresponds to a prograde rotation relative to the projected rotation of the bulge of MKN 421. This feature of the system may be important to the stability of the companion. Future simulations should look in more detail at the effects of retrograde vs. prograde rotation in a tidal encounter. 5. Conclusions We have gathered evidence that supports the following conclusions: 1. MKN 421's nearest companion galaxy MKN 421-5 appears to be an early-type spiral, rather than elliptical. Spectral evidence suggests moderately recent star-formation activity in the companion, a feature which is known to accompany galaxy interactions. 2. MKN 421-5 contains a Seyfert-like nucleus, but without detectable emission lines in ground -- based spectroscopy. We find that its luminosity is probably too high for a compact nuclear star cluster and conclude it is most likely to be an AGN, in spite of the lack of spectroscopic evidence. 3. We confirm the published radial velocity for MKN 421-5. We find that MKN 421-5 is very likely to be bound to MKN 421, and its orbital velocity is consistent with the trend in the bulk rotational velocity of MKN 421, suggesting tidal interaction. 4. We report the first spectroscopic observation of emission lines from MKN 421's nucleus, notably Hα and NII. Both broad and narrow components of Hα are present, with the broad line emission showing a velocity dispersion of several thousand km s−1, typical of QSO emission lines. Spectroscopic monitoring of these lines should continue. 5. The observed orbital velocity of the companion MKN 421-5 provides a lower limit on the mass of MKN 421's bulge of 5.9 × 1011 solar masses, with an estimated bulge mass -- to -- light ratio of ≥ 17. -- 26 -- We thank R. Linfield and the staff of Palomar Observatory for their invaluable help with the observations, and Glenn Piner for his helpful comments on the manuscript. This work was performed at the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The National Radio Astronomy Observatory is a facility of the National Science Foundation, operated under a cooperative agreement by Associated Universities Inc. This research is based in part on observations made with the NASA/ESA Hubble Space Telescope, obtained from the data archive at the Space Telescope Science Institute. STScI is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract NAS 5-26555. REFERENCES Buckley et al. 1996 ApJ Letters 472,9. Comastri, A. F. et al. 1997, ApJ 480, 534. Corbett, E. A., Robinson, A., Axon, D. J., Hough, J. H., Jeffries, R. D., Thurston, M. R., & Young, S. 1996, MNRAS, 281, 737 Falomo, R., 1996, MNRAS 283, 241. Gaidos, J. et al. 1996, Nature 383, 319. Heidt, J. 1999, in Takalo L. O., and Sillanpaa, A. (eds.), Proc. BL Lac Phenomena, PASP159, 367. Hernquist, L., and Mihos, J. C., 1995, ApJ 448, 41. Hickson, P., Fahlman, G. G., Auman, J. R., Walker, G. A. H., Menon, T. K., and Ninkov, Z., 1982, ApJ 258, 53. Holtzman, J. A., et al., 1995, PASP107, 1065. Honma, M., 1999, LANL preprint astro-ph 9904079. Kaspi, S. 1997, in Emission Lines in Active Galaxies: New Methods & Techniques, ed. B.M. Petersen, F.-Z. Cheng, and A.S. Wilson, (San Francisco: ASP), 159. Kikuchi, S., and Mikami, Y., 1987, ApJ 39, 237. Knapp, G.R., Bies, W.I., and van Gorkom, J. H., 1990 AJ99, 476. -- 27 -- Liu, F. K., et al. 1997, A&A Suppl. 123, 569. Laor, A., 1998 ApJ505, L83. Magorrian, J., et al. 1998 AJ 115, 2285. Marcha, M. J. M., Browne, I. W. A., Impey, C. D., & Smith, P. S. 1996, MNRAS, 281, 425 Mukherjee, R., et al., 1997, ApJ 490, 116. Morganti, R., Ulrich, M. -- H., & Tadhunter, C. N. 1992, MNRAS, 254, 546 Oke, J. B. 1990, AJ, 99, 1621 Piner, B. G., Unwin, S. C., Wehrle, A. E., Edwards, P. G., Fey, A. L., and Kingham, K. A., 1999, ApJ, in press. Scarpa, R., Falomo, R., & Pian, E. 1995, A&A, 303, 730 Schmitt, H. R., Kinney, A. M., Storchi-Bergmann, T., and Antonucci, R., 1997 ApJ477, 623. Schweizer, L.Y., 1987 ApJS64, 427. Sitko, M. L., & Junkkarinen, V. T. 1985, PASP, 97, 1158 Shlosman, I., Begelman, M. C., Frank, J., 1990 Nature 345, 679. Simien, F., and Prugniel, Ph., 1997, 122, 521; also 126, 15, 519. Simien, F., and Prugniel, Ph., 1998, 131, 287. Tosti, G., Fiorucci, M., Luciani, M. et al. 1998 A&A339, 41. Ulrich, M. -- H. 1978, ApJ, 222, L3 Ulrich, M. -- H. 1981, A&A, 103, L1 Ulrich, M. -- H., Kinman, T. D., Lynds, C. R., Rieke, G. J., & Ekers, R. D. 1975, ApJ, 198, 261 Ulvestad, J.S, and Wilson, A. S., 1989, ApJ343, 659. van Gorkom, J.H., Knapp, G. R., Ekers, R.D., et al., 1989 AJ97, 708. Veilleux, S., & Osterbrock, D. E. 1987, ApJS, 63, 295 -- 28 -- Vermeulen, R. C., Ogle, P. M., Tran, H. D., Browne, I. W. A., Cohen, M. H., Readhead, A. C. S., & Taylor, G. B. 1995, ApJ, L5 Wandel, A., 1999, Ap. J. Lett. (in press); see also LANL preprint astro-ph 9904370. Wandel,A., Peterson, B., and Malkan, M., 1999, ApJ(in press), (LANL preprint astro-ph 9905224). Zirbel, E. L. 1997, ApJ, 476, 489 Zhang, F. J., and Baath, L. B., 1990, A&A236, 47. This preprint was prepared with the AAS LATEX macros v4.0.
astro-ph/0510208
1
0510
2005-10-07T08:08:21
Testing for w<-1 in the Solar System
[ "astro-ph", "gr-qc", "hep-ph", "hep-th" ]
In scalar-tensor theories of gravity, the equation of state of dark energy, w, can become smaller than -1 without violating any energy condition. The value of w today is tied to the level of deviations from general relativity which, in turn, is constrained by solar system and pulsars timing experiments. The conditions on these local constraints for w to be significantly less than -1 are established. It is demonstrated that this requires to consider theories that differ from the Jordan-Fierz-Brans-Dicke theory and that involve either a steep coupling function or a steep potential. It is also shown how a robust measurement of w could probe scalar-tensor theories.
astro-ph
astro-ph
Testing for w < −1 in the Solar System J´erome Martin,∗ Carlo Schimd,† and Jean-Philippe Uzan‡ Institut d'Astrophysique de Paris, GRεCO-CNRS, UMR 7095, Universit´e Pierre et Marie Curie, 98bis boulevard Arago, 75014 Paris, France (Dated: November 12, 2018) In scalar-tensor theories of gravity, the equation of state of dark energy, w, can become smaller than −1 without violating any energy condition. The value of w today is tied to the level of deviations from general relativity which, in turn, is constrained by solar system and pulsars timing experiments. The conditions on these local constraints for w to be significantly less than −1 are established. It is demonstrated that this requires to consider theories that differ from the Jordan- Fierz-Brans-Dicke theory and that involve either a steep coupling function or a steep potential. It is also shown how a robust measurement of w could probe scalar-tensor theories. 5 0 0 2 t c O 7 1 v 8 0 2 0 1 5 0 / h p - o r t s a : v i X r a PACS numbers: 98.80.Cq, 05.50+h.98.80.Es Various observations indicate that the expansion of our Universe is presently accelerated [1]. While still debated, this conclusion appears to be more and more robust and, as a consequence, the discussion has now mainly shifted to explaining the cause of this acceleration. The property of the effective equation of state inferred from the obser- vations, w, is a key issue in this investigation. In particu- lar, showing that w 6= −1 and/or dw/dz 6= 0 (z being the redshift) would exclude a cosmological constant, proba- bly the most natural candidate and, hence, would have drastic implications for fundamental physics. Recently, various observations have pointed towards the conclu- sion that w < −1 [2]. Although far from being settled, this would also have important consequences since such an equation of state cannot be achieved by quintessence models [3] for which −1 ≤ w ≤ 1. "Phantom models" [4], consisting in a scalar field with a minus sign in front of the kinetic term, are very often advocated in order to explain w < −1. These models are plagued by various theoret- ical problems such as, for instance, their stability when interactions with other fields are taken into account. However, another route can be investigated since theo- ries where the gravity sector is modified, i.e. where grav- ity is no longer described by general relativity (GR), can also entertain w < −1 even if the matter sector is de- scribed in a standard fashion. The prototype of such a theory is a scalar-tensor theory of gravity which is both well-defined and well-motivated [5] as they arise as the low energy limit of string theory. It is worth recalling that they are equivalent to theories where the gravita- tional action is given by an arbitrary function of the Ricci scalar and also encompass the Jordan-Fierz-Brans-Dicke (JFBD) theory as a particular case. As already men- tioned, having 1 + w negative is linked to the fact that the gravity sector is modified but such modifications are strongly constrained by solar system and pulsars timing ∗Electronic address: [email protected] †Electronic address: [email protected] ‡Electronic address: [email protected] experiments. Therefore, one can hope to use these local tests to track the real nature of the dark sector. In this letter, we investigate these issues in general scalar-tensor theories. In the Jordan frame, i.e. in the (physical) frame where the experimental data have their usual interpretation, scalar-tensor theories are described by the action [6] S = 1 16πG∗ Z d4x√−g(cid:2)F (ϕ) R − Z (ϕ) gµν∂µϕ∂νϕ −2U (ϕ)(cid:3) + Sm [Ψm, gµν ] , (1) which depends on three arbitrary functions F (ϕ), Z (ϕ) and the potential U (ϕ), only two of which are indepen- dent. G∗ is a constant, different from the gravitational constant measured in a Cavendish experiment, Gcav, as will be discussed in more details below. In the following, the matter action, Sm [Ψm, gµν ], describes a pressure-less perfect fluid. In the particular case of a Friedmann-Lemaitre- Roberston-Walker (FLRW) Universe, choosing Z (ϕ) = 1, the field equations reduce to 3(cid:18)H 2 + K −(cid:18)2 a2(cid:19) = 8πG∗ a2(cid:19) = + H 2 + K a a ρm F ϕ2 2F + ϕ2 2F − 3H F F + 2H + F F F F − U F U F , (2) , (3) + with H ≡ a/a, a being the scale factor and K the curvature of the spatial sections. The energy density of matter scales as a−3. These equations should be compared to 3(H 2 + K/a2) = 8πGcav(ρm + ρDE ) and −(cid:0)2a/a + H 2 + K/a2(cid:1) = 8πGcavpDE, since the equa- tion of state of dark energy is experimentally inferred from the expansion history of the Universe by using the standard Friedmann equations of GR. It follows that 3ΩDEw ≡ −1 + ΩK + 2q where q ≡ −aa/( a)2 is the acceleration parameter, ΩK ≡ −K/a2H 2 and that the dark energy density parameter is defined as ΩDE (z) ≡ K(1 + z)2, where here and in the H 2/H 2 rest of this letter the subscript "0" denotes the present day value of the corresponding quantity. Clearly, iden- tifying ρDE and pDE from Eqs. (2) and (3) leads to the m(1 + z)3− Ω0 0 − Ω0 conclusion that, in scalar-tensor theories, w needs not to be positive definite, depending on the choice of F . This was already stressed by various works [7] and worked out in the particular case of JFBD theories in Refs. [8]. The present values of F and its derivatives are con- strained by the solar system and pulsars experiments. In order to reveal the link between w and these con- straints, it is more convenient to work in the Einstein frame. The Einstein frame metric, g∗ µν (all Einstein frame quantities will be denoted with a star), is related to the physical metric through the conformal transformation µν = F (ϕ) gµν. Setting F (ϕ) = A−2(ϕ∗), the scale g∗ factors and cosmic times in both frames are related by a = Aa∗ and dt = Adt∗, so that the Hubble parameters are linked by AH = H∗ (1 + αϕ′ ∗) with H∗ = d ln a∗/dt∗ and where a prime denotes a derivative with respect to the Einstein frame number of e-folds, N∗ ≡ ln(a∗/a∗0). Deviations from GR are usually described by two pa- rameters, α and β, which are the first and second field derivative of the coupling function A(ϕ∗), namely α ≡ d ln A dϕ∗ , β ≡ dα dϕ∗ . (4) In this framework, GR is characterized by α = β = 0 while the JFBD theory corresponds to a constant α2 ≡ (2ωBD + 3)−1 and β = 0. As mentioned before, in scalar-tensor theories, the Newton constant obtained in a Cavendish experiment differs from G∗ and is given by 2 Gcav = G∗A2 0(1 + α2 0) . (5) 0/(1 + α2 The parameters α0 and β0 are constrained by various ex- periments. If we define ¯γ ≡ γ PPN − 1 and ¯β ≡ β PPN − 1, where γ PPN and β PPN are the post-Newtonian parame- ters [9], related to α0 and β0 by ¯γ = −2α2 0) and 2 ¯β = β0α2 0)2, then the perihelion shift of Mer- cury implies (cid:12)(cid:12)2¯γ − ¯β(cid:12)(cid:12) < 3 × 10−3 while the Lunar Laser Ranging experiment sets 4 ¯β − ¯γ = −(0.7 ± 1) × 10−3. On the other hand, a bound on ¯γ alone is set from the time delay variation to the Cassini spacecraft near solar conjunction, namely ¯γ = (2.1 ± 2.3) × 10−5, see Ref. [10] for a review. We conclude that 0/(1 + α2 α2 0 < 10−5, β0 & −4.5 , (6) where the lower bound on β0 arises from pulsar timing experiments [10]. Note, however, that we cannot con- sider arbitrarily large values of β0 since then the post- Newtonian approximation scheme would breakdown. In this case, the upper constraints should be reanalyzed. We thus loosely assume that β0 . 100. Working out the Friedmann equations in Einstein frame [6], we obtain that the equation of state is given by 3ΩDEw = −1 + ΩK + (1 + αϕ′ ∗)(cid:20)3 A2 0) − 2(cid:21) Ωm − 2(1 + αϕ′ A2 0(1 + α2 ∗)ΩDE + 2 αϕ′ ∗(2 + αϕ′ 1 + αϕ′ ∗ ∗) + ϕ′2 ∗ ∗ + βϕ′2 ∗ αϕ′′ (1 + αϕ′ ∗)2 , (7) − 2 where Ωm ≡ 8πGcavρm/3H 2 so that ΩDE + Ωm + ΩK = 1. In the limit of a minimally coupled scalar field, it reduces to the standard relation 1 + w = 2ϕ′2 ∗ /3ΩDE. At this point, it is of utmost importance to stress that ∗ is not free. Indeed, in the Einstein frame, the value of ϕ′ the Friedmann equation reads [6] H 2 ∗ (cid:0)3 − ϕ′2 ∗ (cid:1) = −3 K a2 ∗ + 8πG∗ρm∗ + 2V (ϕ∗) , (8) where V (ϕ∗) = U/(2F 2), and the positivity of the energy ∗ < √3, as long as density of matter implies that ϕ′ ΩK ≪ Ωm (in the following, we assume that ΩK = 0). The expression (7) for w is completely general but very intricate and hence not so illuminating. However, taking into account that α0 has to be small and that ϕ′ ∗ has to be bounded by √3 and, therefore, that α0ϕ′ ∗0 has to be small as well, the present day value of the equation of state simplifies considerably and reduces to ∗0 − 2α0ϕ′′ ∗0. DE(1 + w0) ≃ 2(1 − β0)ϕ′2 3Ω0 (9) This formula turns out to be our main result. The contri- bution of β0 arises from the term F /F in the right hand ∗0 positive and large compared to ϕ′2 side of Eq. (3). In fact, Eq. (9) shows that w0 < −1 is always possible provided β0 > 1 (in this case, even if ∗0 ≪ ϕ′ the slow-roll approximation is satisfied, i.e. ϕ′′ ∗0) and/or αϕ′′ ∗0. Both regimes cannot be reached in the case of JFBD theories (except if the time variation of Gcav is large, see below) and exist even in the limit α0 → 0 so that all local tests can be satisfied. The amplitude of w0 depends on the value of ϕ′ ∗0 and ϕ′′ ∗0. Independently of any dynamics, these two quantities are constrained by the bounds on the time variation of the gravitational constant [11]. Defining d ln Gcav dt ≡ σH , d2 ln Gcav dt2 ≡ ξH 2 , (10) the parameter σ0 is bounded by σ0 < 5.86 × 10−2h−1. There is no stringent bounds on ξ0 but we can estimate that, since σ0 has been "measured" during a period of 0 . G/G0/(20 yr). This about 20 years, we have ξ0H 2 implies that ξ0 . 5 × 108h−1σ0 ∼ 2.5 × 107h−2. Using Eq. (5), one can then express ϕ′ ∗ and ϕ′′ ∗ in terms of the parameters σ and ξ. As long as β 6= −(1 + α2), a case we shall discuss later, one arrives at while the second derivative reads ϕ′ ∗ = σ 2α (cid:18)1 + β 1 + α2 − σ 2(cid:19)−1 , (11) ϕ′′ ∗ = (1 + αϕ′ ∗)2 2α [1 + β/ (1 + α2)] ξ − ϕ′ ∗)2(cid:20)ΩDE −(cid:18)3 ∗(cid:26)ϕ′ A2/A2 1 + α2 ∗(cid:20) β α − α + 0 − 2(cid:19) Ωm 0 +(1 + αϕ′ 1 1 + α2 + β (cid:18) dβ dϕ∗ − 2αβ 1 + α2(cid:19)(cid:21) 2 − 1(cid:21) − ϕ′2 ∗ (cid:27). 3 (12) Interestingly enough, we see that ϕ′′ ∗0, and hence w0, de- pends on ξ0 and on the derivative of the β function. The latter is not constrained and we will assume that dβ/dϕ∗ . O(100) for the same reason as for β. The above formula is rather complicated but, given the pre- vious constraints, we simply have αϕ′′ ∗ ≃ ξ 2 + 2β/ (1 + α2) − ×(cid:18)1 + 1 + α2 β 1 − α2 1 + α2(cid:19) ϕ′2 ∗ . β 1 + β/ (1 + α2) (13) In particular, the unknown term dβ/dϕ∗ does not appear in this approximation. As mentioned above, the previous considerations are independent from the dynamics. However, from a model building point of view, ϕ′′ ∗ are not independent once the potential V (ϕ∗) is chosen. They are related through the Klein-Gordon equation which reads ∗ and ϕ′ 2(X + 1) 3 − ϕ′2 ∗ ϕ′′ ∗ + (2 + X)ϕ′ ∗ = −(αX + αv) , (14) with X ≡ Ωm/ΩU , where ρU ≡ 2U/(16πG∗) and αv ≡ d ln V /dϕ∗. When the kinetic energy of the scalar field is negligible, we have ΩU ∼ ΩDE . We now come back to our master equation (9) and an- alyze the two regimes where 1 + w0 can become negative and even large (in absolute value). The first regime is the natural extension of quintessence models to scalar- tensor theories [12] and corresponds to the situation in which the field is decelerating, i.e. ϕ′′ ∗0, so that α0ϕ′′ ∗0 is negligible in Eq. (9). Clearly, this requires β0 > 1. Such a regime, that cannot be reached in a JFBD model, has the advantage to exhibit an attraction mechanism toward GR [13] so that α0 can dynamically be made small. Since, in this situation, β0 is not close to −1, Eq. (11) implies that ϕ′ ∗0 ∼ σ0/[2α0(1 + β0)] and Eq. (13) leads to ξ0 ∼ β0σ2 0/[2α2 0(1 + β0)]. On the other hand, from the Klein-Gordon equation, one obtains that ϕ′ ∗0 ∼ −(α0X0 + αv0)/(2 + X0) and, in order to fulfill the ∗0 < √3, one must have αv0 . 4.2. The two condition ϕ′ ∗0 ≪ ϕ′ expressions for ϕ′ if we insert the result in Eq. (9), one gets ∗0 can be used to infer what β0 is and, 3Ω0 DE(1 + w0) ≃ σ0αv0 (2 + X0)α0 < 0 , (15) in the most interesting limit where αv0 ≫ α0X0 since, otherwise, 1 + w0 ∼ O(α2 0). They are two ways to inter- pret Eq. (15). Either it gives w0 in terms of α0 and αv0, assuming that σ0 is known. Or it provides, for fixed α0 and αv0, the minimum value that w0 can reach assum- ing, as is the case now, that only an upper bound on σ0 is available. Fig. 1 depicts the above-mentioned minimum value as a function of (α0, αv0). At this point several remarks are in order. Firstly, it is unclear whether the above situation can be real- ized without some fine-tuning in a realistic model, where both the coupling function and the potential are running away in order to have attraction towards GR and insen- sitivity to the initial conditions [14]. Let us illustrate this point on the following particular example. Consider the case where α ∝ e−λϕ∗ so that β0 = −α0λ. Since we need α0 ≪ 1 and, at the same time, β0 > 1, this means that λ ≫ 1 which may be considered as unnatu- ral. Secondly, one can reverse the logics and study what a detection of w0 < −1 would imply on scalar-tensor theories (assuming, of course, that a slow-rolling ϕ∗ is causing the acceleration of the expansion). Besides the fact that the condition β0 > 1 would drastically improve the current limit on β0 and, in particular, exclude GR, it is also interesting to remark that this would link w0 to the time variation of the Newton constant through the expressions σ2 0/(1 − β0) DEβ0(β0 + 1)(w0 + 1)/(1 − β0) . 6 × 102, a and ξ0 ∼ 3Ω0 limit sharper than the bound set by local experiments. Finally, it is interesting to notice that w0 < −1 can be obtained even if the potential energy is negligible (in GR In this case, one this would correspond to w0 ∼ 1). can show that ϕ′ ∗0 ∼ −3Ω0 ∗0/2 so that α0ϕ′′ ∗0 is indeed negligible in Eq. (9). Then, the equation of state can easily be obtained and reads 1 + w0 ∼ 2(1 − β0). Again β0 is bounded by the con- straint on σ0, 1 + β0 < σ0/(α0p12Ω0 DE), so that a large DE(1 + w0)(1 + β0)2α2 ∗0 ∼ (3Ω0 DE)1/2 and ϕ′′ 0 ≃ 6Ω0 mϕ′ -1 -1.5 -1.5 w 0 min -2 -2 -2.5 -2.5 -3 -3 -3.4 -3.6 0 0 -1 -1 -3.8 log10 Α0 -4 -4 -4 -3 -3 -2 -2 ΑV0 ΑV0 FIG. 1: Minimum value of w0 as a function of (α0, αv0) for σ0 < 10−3. value of w0 requires σ0 ≪ α0. The second regime corresponds to the case where αϕ′′ ∗ dominates in Eq. (9). As can be seen in Eq. (13), this is possible if ξ0 is large and positive and/or if β0 ∼ −1. Strictly speaking, one should in fact consider the limit β0 → −(1 + α2 0) corresponding, when V = 0, to the Barker theory [15] in which A = cos ϕ∗ and Gcav con- ∗ and ϕ′′ stant (σ = ξ = 0, whatever the value of ϕ′ ∗ ). Since α = − tan ϕ∗, the cosmological evolution drives the theory away from GR unless a potential keeps ϕ∗ close to 0 until the last e-folds inducing a large variation of αϕ′′ ∗ ∼ O(1) in order to have 1 + w0 < 0. Such a model seems very contrived unless the potential exhibits a slope discontinuity recently. 0 . 2 × 10−3. 0, Eq. (13) leads to αϕ′′ 0β 2 Let us now assume that ξ0 ∼ 0 since, from dimen- In ∗0 ∼ 0) and w0 diverges when β0 → ∗ < √3 implies that sional analysis, one expects ξ0 ∼ σ2 the limit β0 → −1 − α2 0 ϕ′2 2α2 −1 − α2 ∗0/(1 + β0 + α2 0. However, the constraint ϕ′ 1 + β0 + α2 4 > σ0/(2√3α0) and, hence, the smallest value 0 ∼ that can be obtained is (1 + w0)min ∼ −8√3α3 DEσ0). In particular, w0 ≪ −1 is perfectly possible in this regime if σ0 ≪ α3 0. Finally, let us see what this regime implies in terms of model building. The Klein-Gordon equation implies that 2ϕ′′ ∗0/3 ∼ −αv0/(2 + X0) so that the slope of the potential must be very large, αv0 ≫ α−1 0 . Moreover since, in this regime, β0 < 0 the theory is driven away from GR and it follows that the potential must be tuned in order to prevent this drift. 0/(Ω0 As a conclusion, let us summarize our main findings. Firstly, we have confirmed that 1 + w0 is not positive definite in scalar-tensor theories even if all the matter energy conditions are satisfied. Secondly, we have estab- lished under which conditions 1+w0 can become negative given the local constraints coming from solar system and pulsars measurements (in the case of chameleon mod- els [16], it has been argued that α0 can be of order unity today, a case where large negative values of 1 + w0 can be achieved more easily). We have shown that getting a non negligible deviation from −1 necessarily implies a non vanishing β0 (except if ξ0 is large), a situation that cannot be reached in the JFBD case. Thirdly, in terms of model building, we have demonstrated that this cor- responds either to β0 > 1, a situation where the scalar field is slow-rolling today and the coupling constant is very steep or to β0 ∼ −1 (or ξ0 large) where the slope of the potential is very large. Fourthly, we have also shown how a measurement of w0 < −1 could improve the local constraints on the deviations from GR. This highlights the complementarity [17] between cosmological and lo- cal tests of gravity. Finally, let us remark that we have only taken into account the local constraints. A next step would be to reconstruct the redshift evolution of the models and to show that they are not pathological [6]. Acknowledgments: we thank G. Esposito-Far`ese for many enlightening comments and discussions. [1] S. Perlmutter et al. , Nature (London) 391, 51 (1998); S. Perlmutter et al. , Astrophys. J. 517, 565 (1999); P. M. Garnavich et al. , Astrophys. J. 493, L53 (1998); A. G. Riess et al. , Astron. J. 116, 1009 (1998). [2] H. K. Jassal, J. S. Bagla and T. Padmanabhan, astro-ph/0506748. [3] B. Ratra and P. J. E. Peebles, Phys. Rev. D 37, 3406 (1988); C. Wetterich, Astron. Astrophys. 301, 321 (1995); P. G. Ferreira and M. Joyce, Phys. Rev. D 58, 023503 (1998); P. Brax and J. Martin, Phys. Lett. 468B, 40 (1999); P. Brax and J. Martin, Phys. Rev. D 61, 103502 (2000); P. Brax, J. Martin and A. Riazuelo, Phys. Rev. D 62, 103505 (2000). [4] R. Caldwell, Phys. Lett. B 545, 23 (2002). [5] T. Damour and G. Esposito-Far`ese, Class. Quant. Grav. 9, 2093 (1992). A. Riazuelo and J. P. Uzan, Phys. Rev. D 66, 023525 (2002). [8] D. Torres, Phys. Rev. D 66, 043522 (2002); S. M. Carroll, A. De Felice and M. Trodden, Phys. Rev. D 71, 023525 (2005). [9] C. Will, Theory and Experiments in Gravitational Physics, Cambridge University Press, 1993. [10] G. Esposito-Far`ese, gr-qc/0409081. [11] J. P. Uzan, Rev. Mod. Phys. 75, 403 (2003). [12] J. P. Uzan, Phys. Rev. D 59, 123510 (1999). [13] T. Damour and K. Nordtvedt, Phys. Rev. Lett. 70, 2217 (1993). [14] N. Bartolo and M. Pietroni, Phys. Rev. D 61, 023518 (1999). [15] J. Barker, Astrophys. J. 219, 5 (1978). [16] J. Khoury and A. Weltman, Phys. Rev. Lett. 92, 171104 [6] G. Esposito-Far`ese and D. Polarski, Phys. Rev. D 63, (2004). 063504 (2001). [17] J. P. Uzan, N. Aghanim, and Y. Mellier, Phys. Rev. D [7] B. Boisseau et al., Phys. Rev. Lett. 85, 2236 (2000); 70, 083533 (2004); J. P. Uzan, astro-ph/0409424.
astro-ph/0301140
1
0301
2003-01-08T19:27:18
The calibration of interferometric visibilities obtained with single-mode optical interferometers. Computation of error bars and correlations
[ "astro-ph" ]
I present in this paper a method to calibrate data obtained from optical and infrared interferometers. I show that correlated noises and errors need to be taken into account for a very good estimate of individual error bars but also when model fitting the data to derive meaningful model parameters whose accuracies are not overestimated. It is also shown that under conditions of high correlated noise, faint structures of the source can be detected. This point is important to define strategies of calibration for difficult programs such as exoplanet detection. The limits of validity of the assumptions on the noise statistics are discussed.
astro-ph
astro-ph
Astronomy & Astrophysics manuscript no. (DOI: will be inserted by hand later) November 8, 2018 3 0 0 2 n a J 8 1 v 0 4 1 1 0 3 0 / h p - o r t s a : v i X r a The calibration of interferometric visibilities obtained with single-mode optical interferometers Computation of error bars and correlations G. Perrin1 LESIA, FRE 2461, Observatoire de Paris, section de Meudon, 5, place Jules Janssen 92190 Meudon, France e-mail: [email protected] Received 4 September 2002 / Accepted 19 December 2002 Abstract. I present in this paper a method to calibrate data obtained from optical and infrared interferometers. I show that correlated noises and errors need to be taken into account for a very good estimate of individual error bars but also when model fitting the data to derive meaningful model parameters whose accuracies are not overestimated. It is also shown that under conditions of high correlated noise, faint structures of the source can be detected. This point is important to define strategies of calibration for difficult programs such as exoplanet detection. The limits of validity of the assumptions on the noise statistics are discussed. Key words. techniques: interferometric -- methods: data reduction 1. Introduction 2. Principles of data reduction and calibration With optical-infrared interferometry becoming more ma- ture, the quality of visibility measurements have become an issue. Single-mode interferometers (see Sect. 2.3) allow one to eliminate non-stationary effects by filtering out the spatial modes of turbulence. The response of interferom- eters is therefore very stable and the issue of estimating the accuracies of non-biased data is raised. The final visi- bility estimate is a complex quantity as it is a non-linear mix of noisy measurements and of parameter estimates with their own uncertainties. Estimating the stability of the instrument, a crucial point for calibration, and the fi- nal error on visibilities is therefore non-trivial and must be considered with caution. Moreover, data analysis mainly consists of model fitting the final visibilities and the mat- ter of their potential correlations becomes important, es- pecially if some very faint structures are looked for, as is the case in extra-solar planet detection. In this paper I propose a method to meet these challenges. The method has been tested and elaborated along with the FLUOR interferometer, the first single-mode interferome- ter. This method was first published in Perrin (1996) and used in Perrin et al. (1998). It is updated and improved in this paper by accounting for correlations. Send offprint requests to: G. Perrin In this section the general scheme of data reduction is reviewed to introduce the vocabulary and notations. Two main steps are to be considered. In the first one (Sect. 2.1), the fringe processing, fringe contrasts are derived from raw signals. Because of contrast losses, fringe contrasts are cal- ibrated in a second step (Sect. 2.2) to provide the visibil- ities directly linked to the spatial intensity distribution of the source. 2.1. Fringe contrasts estimates In the following we will distinguish between the fringe con- trast obtained on a source and the visibility of this source. The fringe contrast measured from a single exposure or scan is called the coherence factor and is noted µ whereas the visibility is noted V . Whatever the beamcombining technique, µ being the modulus of a complex number, unbiased estimators are only obtained for squared quantities from wich biases due to additive noise can be subtracted. In the future, phase referencing techniques may allow one to directly measure complex visibilities (real and imaginary parts) but this is not the case yet and I will only consider measurements of fringe contrast moduli. The result of the processing of a series of scans on a single source is a series of realizations of the µ2 estimator or is an average value of the realiza- 2 G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers t n u o C 30 25 20 15 10 5 0 71 UMa <m 2> = 0.8022 rms = 0.1474 r = 0.67 t n u o C 30 25 20 15 10 5 0 Sge <m 2> = 0.9842 rms = 0.1301 r = 0.57 t n u o C 35 30 25 20 15 10 5 0 V636 Her <m 2> = 0.7268 rms = 0.0975 r = 0.27 0,4 0,5 0,6 0,7 0,8 m 2 0,9 1 1,1 1,2 0,6 0,7 0,8 0,9 1 m 2 1,1 1,2 1,3 1,4 0,4 0,5 0,6 0,7 0,8 0,9 1 1,1 m 2 Fig. 1. Examples of squared coherence factor histograms obtained with FLUOR in one of the interferometric channels. About 100 interferograms have been recorded for each object. The mean and rms of individual measurements are given for this channel. The correlation factor r measures the noise correlation between the two interferometric channels. The amount of atmospheric piston is decreasing from the left to the right. tions with a 1σ error bar if their statistical distribution can be trusted to be Gaussian. 2.2. Necessity for a calibration The average µ2 is not directly an estimator of the squared modulus of the visibility of the source because some physi- cal phenomena degrade the coherence factor. Among these phenomena, polarization mismatches between the inter- ferometric beams are the most common. Without perfect beam cleaning by a fiber, atmospheric turbulence also de- grades the fringe contrast. It is necessary to estimate the loss of coherence on a calibrator source for which the visi- bility is known. A transfer function T is obtained by com- puting the ratio of the measured coherence factor µc to the expected visibility Vexp. With squared quantities this yields: T 2 = µ2 c V 2 exp . (1) If the instrument is stable enough then the estimate of T 2 obtained on the calibrator can be used to derive an estimated visibility for the source from the measured co- herence factor: V 2 = µ2 T 2 = coT 2 µ2. (2) where I call coT 2 the squared co-transfer function. Its use will be detailed in Sect. 4. 2.3. Assumptions in the case of a single-mode interferometer The calibration process may fail if the assumption that the transfer function is stable is wrong. This usually hap- pens if non-stationary processes like atmospheric phase turbulence play a role in the fringe formation process. In a perfect single-mode interferometer in which single-mode fibers are used to filter the phase aberrations produced by atmospheric turbulence (except for the piston mode) these non-stationary effects are eliminated. In order to avoid in- stabilities due to the piston mode, the fringes are scanned at a frequency far above the characteristic frequency of piston. In interferometers where the piston is stabilized by a fringe tracking servo loop, this issue is solved. The remaining main sources of variation of the transfer func- tion are basically temperature drifts and differential po- larisation effects due to the change of beam inclination on the first mirrors with changing positions of the sources in the sky. In both cases the transfer function drifts are very slow and a good estimate of the transfer function can be obtained by interpolating two estimates bracketing the source to be calibrated. This has been demonstrated with the FLUOR beamcombiner, as will be shown is Sect. 7.2. In the following we will therefore consider that the effi- ciency of the interferometer is continuously assessed by observing calibrators before and after science sources. We will not consider the case where the transfer function is derived by averaging individual transfer functions on a large temporal scale, as this is not required for a single- mode interferometer. This technique does not allow one to assess the quality of the calibration in detail. 3. Estimating fringe contrasts This section focuses on estimating the statistical prop- erties of fringe contrasts. I will not describe the method to compute coherence factors from single exposures and I will refer the reader to appropriate articles in the next paragraphs. d G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers 3 May 15, 2000 1,8 1,6 1,4 1,2 1 1,8 1,6 1,4 1,2 1 2 T o c 2 T o c 6:20 6:40 7:00 7:20 7:40 8:00 8:20 8:40 9:00 UT Time (hh:mm) May 22, 2000 5:50 6:40 7:30 8:20 9:10 10:00 10:50 11:40 12:30 UT Time (hh:mm) Fig. 2. Examples of squared co-transfer functions measured with FLUOR. The two curves for each night correspond to the two interferometric channels of the coaxial interferometer. The full circles are the squared co-transfer functions measured on calibrators whereas the open circles are the values interpolated at the time when the science targets were observed. 1 σ error bars are displayed. 3.1. Single-channel spatial modulation interferometer The estimate of the variance of the coherence factor esti- mator µ2 is then: In a multiaxial interferometer, distant parallel beams feed a focusing optic. The beams are recombined in the focal plane where they overlap at the focus locus. The modula- tion is spatial as the fringe phase varies across the diffrac- tion pattern. A method to derive fringe contrasts has been published by Mourard et al. (1994) in the case of GI2T. The method has been adapted to AMBER which is a single-mode multiaxial interferometer (Chelli et al. 2000). Thanks to the filtering of the non-stationary modes of tur- bulence, the statistics of µ2 can be well approximated by a Gaussian distribution. This will be demonstrated in the case of the data obtained with FLUOR in Sect. 7.1. The estimate of the squared coherence factor is therefore the mean of the distribution of the realizations denoted µ2. An unbiased estimate of the variance of individual mea- surements is: S2 = 1 N − 1 N −1 (µ2 n − µ2). Xn=0 (3) V ar(µ2) = S2 N . (4) 3.2. Two-channel temporal modulation interferometer In a coaxial interferometer, beams are superimposed in po- sition and in direction. This can be realized with a beam- splitter or with a fiber coupler. A relative phase between the beams is introduced by setting an optical path differ- ence. This is achieved with a moving mirror in one of the two beams, hence the temporal modulation of the phase. A method to compute fringe contrasts for this type of inter- ferometer is described in Coud´e du Foresto et al. (1997). A more recent method based on wavelets analysis has been proposed by Segransan et al. (1999). The method to ob- tain an estimate of the coherence factor without the pho- ton noise bias is explained in Perrin (2002). A prototype instrument for this kind of interferometer is the FLUOR beamcombiner. The difference with the previous interferometer of Sect. 3.1 is that it produces two interferometric beams and there- fore two sets of coherence factors estimates. The statistics 4 G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers y t i l i b i s i V 0,75 0,7 0,65 0,6 0,55 0,5 1 0,8 0,6 0,4 0,2 0 0 Without correlations =17,05±0,13 mas UD c 2=0,92 With correlations =17,27±0,31 mas UD c 2=13,30 3 0 3 2 3 4 3 6 3 8 4 0 0,75 0,7 0,65 0,6 0,55 0,5 3 0 3 2 3 4 3 6 3 8 4 0 1 0 2 0 3 0 4 0 5 0 6 0 Spatial frequency (cycles/arcsec) Fig. 3. Example of visibility fit. The source is SW Vir and the model is a uniform disk. Errors are 1 σ errors. Open circles and dashed line are the visibilities and model fit computed without taking correlations into account. Full circles and continuous line are the equivalent with the method described in this paper. The two cases are separately presented in the little windows. of each set can be well approximated by a gaussian statis- tics as will be shown in Sect. 7. The photon noises of the two interferometric signals are uncorrelated. The read-out noises are generally considered uncorrelated but some cor- relation may occur as different pixels share the same read- out electronics. In addition the two beams suffer from the same turbulence effects (residual piston and photometric beam fluctuations) which generate some noise in the mea- surements. Part of the noise is therefore common to the two signals and the coherence factors estimates are corre- lated. The correlation factor r is directly estimated from the µ2 distributions: r = 1ihµ2 2 − µ2 2i 1)V ar(µ2 2) hµ2 1 − µ2 pV ar(µ2 , (5) where the subscripts describe the two interferometric channels. 4. Estimating the transfer function It is assumed that the transfer function is a slowly varying function which is rapidly sampled. This property will be illustrated with real data in Sect. 7. It is then legitimate to linearly interpolate the squared transfer function at the time when the science source was observed. Because the variances of products of random variables are more eas- ily calculated than those of ratios, the reciprocal of the squared transfer function, the squared co-transfer func- tion, is interpolated instead of the squared transfer func- tion. The use of one or the other is equivalent. In order to be general, two interferometric outputs are always consid- ered. The particular case of the multiaxial interferometer will be considered in discussions. The expression of the in- terpolated co-transfer functions in the two outputs of the instrument is : coT 2 coT 2 1 = x V a 2 µa 2 = x V a 2 µa 2 + (1 − x) V b 2 µb 2 + (1 − x) V b 2 µb 2 1 2 1 2 2   (6) f f G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers 5 X Her G Her y t i l i b i s i V y t i l i b i s i V 1 0,8 0,6 0,4 0,2 0 0 1 0,8 0,6 0,4 0,2 0 0 =12,38±0,04 mas U D c 2=1,48 20 40 60 80 100 Spatial frequency (cycles/arcsec) BK Vir =11,49±0,40 mas U D c 2=0,23 20 40 60 80 100 Spatial Frequency (cycles/arcsec) y t i l i b i s i V y t i l i b i s i V 1 0,8 0,6 0,4 0,2 0 0 1 0,8 0,6 0,4 0,2 0 0 =16,32±0,08 mas U D c 2=1,03 20 40 60 80 100 Spatial Frequency (cycles/arcsec) R Hya =23,75±0,11 mas U D c 2=3,99 20 40 60 80 100 Spatial Frequency (cycles/arcsec) Fig. 4. Examples of fits of visibility data with a uniform disk model. All correlations are taken into account. Errors are 1 σ errors. The visibility point around 85 cycles/arcsec in the G Her panel is not taken into account in the fit. V a and V b are the expected visibilities of calibrators A and B. Coefficient x is the relative time distance between the observation of the science target and the observation of calibrator A. The expected visibilities are supposed to be Gaussian random variables. Dropping the channel indices, the variances of the squared co-transfer functions are equal to: V ar(coT 2) = x2V ar(coT a2) + (1 − x)2V ar(coT b2 (7) ) when the two calibrators are different. When the two cal- ibrators are the same then the variance is equal to: V ar(coT 2) = x2V ar(coT a2) + (1 − x)2V ar(coT b2 (8) ) +2 x(1 − x) µa2µb2 V ar(V a2) The squared co-transfer functions estimated on the cali- brators are ratios of Gaussian estimators. These new ran- dom distributions are not Gaussian. They are Cauchy distributions (the density probability of which is a Lorentzian) with no mean and no variance. By analogy with the standard deviation of a Gaussian law, an esti- mate of the uncertainty can be derived from the width of the confidence interval. The upper and lower limits of the confidence interval at 68.3% for a ratio of two Gaussian random variables α and β are given by (Pelat (1992)): L± = α ± β ∓ β/σβ ((α/σα)2+(β/σβ )2−1)1/2 σα ((α/σα)2+(β/σβ )2−1)1/2 σβ α/σα (9) This interval is not symmetric. I choose as error bar the larger distance of the mean to the limits, thus slightly overestimating the error. The probability that the true value is in this interval around the mean is therefore larger than 68.3%. In the following I will consider that ratios of Gaussian variables are Gaussian variables. This is not rigourously true but this allows one to derive expressions otherwise difficult to handle. The Gaussian and Lorentzian functions are both bell curves, the wings of the latter be- ing more extended than that of the Gaussian. The ap- proximation amounts to giving more weight to the center of the lorentzian. The other possibility would be to apply Monte-Carlo or bootstrap methods to all error bar com- putations, which would make data reduction a very long process for a very limited gain. Nevertheless, this method f f f f 6 G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers will have to be applied once to compute the correlation be- tween the visibilities of a two-channel coaxial interferom- eter (Sect. 5). Consistency of error bars will be addressed in Sect. 7. This final consistency is the justification of the approximations performed. 5. Estimating visibilities 5.1. Mean and variance of channel visibilities The single-channel squared visibility in the case of a mul- tiaxial interferometer or the two-channel squared visibili- ties of a coaxial interferometer are simply the product of the science target squared coherence factors and squared co-transfer functions of Eq. (6) yielding:   V 2 1 = x V 2 2 = x 1 1 1 2 V a2 + (1 − x) µ2 µ2 µa µb µ2 2 V a2 + (1 − x) µ2 µa µb 1 2 2 2 2 2 V b2 2 V b2 (10) The variances of the two-channel squared visibilities are calculated with the variances of the squared co-transfer functions and of the squared coherence factors with the following formula : V ar(AB) = V ar(A)V ar(B) + ¯A2V ar(B) (11) expected visibilities in the two interferometric channels. This can be illustrated by the equation below when the first and second calibrators are different: x2 µ2 µa 1 1 2 ρ12 = 2 2 µ2 2 V ar(V a2) + (1 − x)2 µ2 µa µb 1 )V ar(V 2 2 ) 2 1 1 µ2 2 V ar(V b2 2 µb 2 ) (14) When the two calibrators are the same the correlation fac- tor has the particular expression (for sake of simplicity the two estimates of the visibility at slighlty different baselines are supposed to be the same): pV ar(V 2 2(cid:21)V ar(V a2) (15) ρ12 =(cid:20)x 1 µ2 2 +(1 − x) µ2 µa µb 1 2 2 µ2 2 +(1 − x) µ2 µa µb 1 )V ar(V 2 2 ) 2 1 1 2 2(cid:21)(cid:20)x pV ar(V 2 It is now easy to see from Eq. (10) that if the noise on the measurements is negligible with respect to the uncertain- ties on the expected visibilities of the calibrators then the correlation tends towards 1. In this case, the second in- terferometric channel brings no extra information except a consistency check and the precision on the visibility of the science target is directly proportional to the precision on the expected visibilities of the calibrators. + ¯B2V ar(A) We therefore have: V ar(V 2 V ar(V 2 1 ) = V ar(µ2 +coT 2 2 ) = V ar(µ2 +coT 2 1)V ar(coT 2 1 V ar(µ2 1) 2)V ar(coT 2 2 V ar(µ2 2) 1 ) + µ2 1V ar(coT 2 1 ) 2 ) + µ2 2V ar(coT 2 2 ) 5.3. Comments on visibility variances (12) It is also interesting to analyze the propagation of noises in the visibility estimates. For example, if the noise on the measurements is negligible, it is possible to evaluate the amount of variance due to the uncertainty on the calibra- tors visibilities (or diameters). Dropping channel indices I obtain: 5.2. Correlation between two channel visibilities The above equations define the uncertainties on the chan- nel estimates of the visibilities. In the case of a multiaxial interferometer, this is the final estimate of the visibility. In the case of coaxial interferometers, the two estimates of the visibility need to be averaged at this stage. For this, it is necessary to assess the correlation factor between the two estimates. By definition, the correlation factor is equal to: (13) ρ12 = 1 )(V 2 2 − V 2 2 )i 1 )V ar(V 2 2 ) h(V 2 1 − V 2 pV ar(V 2 This quantity is defined by sums, ratios and products of ten random variables. The correlation factor has to be computed with a Monte-Carlo method by simulating each random variable from its mean and variance (assuming it has Gaussian statistics) and by correlating the series of V 2 2 . In the special case when the correlations between measured quantities can be neglected because the correlated noise level is far below that of the uncorrelated noise, it is to be noticed that there is still a correlation due to the common estimated values of the calibrators 1 and V 2 µa2#2 V ar(V 2)µ,µa,µb = x2" µ2 V ar(V a2) (16) µb2#2 +(1 − x)2" µ2 V ar(V b2 ) If the uncertainties on the calibrators´ squared visibilities are equal to 1% and if the coherence factors are all equal to 1 and observations are equally spaced in time then the uncertainty on the measured squared visibility is equal to 0.7%. This equation also shows that the smaller the visibility of the calibrator, the more amplified the noise on its expected visibility is. Symmetrically, the smaller the visibility of the science target, the smaller the contribution to the noise of the calibrators´ expected visibilities. 5.4. Final estimate of the visibility in a two-channel interferometer The final squared visibility V 2 is estimated from the two squared visibilities obtained from each output of the in- terferometer V 2 2 and their respective variances (or equivalently uncertainties σ1 and σ2). I define the final estimate V 2 as being the least squares fit estimator of 1 and V 2 G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers 7 the squared visibility as this is an optimal estimator for Gaussian random variables. In this fit, the model is lin- ear and has only one parameter: V 2. Let us call C the covariance matrix of V 2 1 and V 2 2 : C =(cid:20) σ2 1 ρ12σ1σ2 ρ12σ1σ2 σ2 2 (cid:21) . (17) The quantity to minimize in the least squares fit can then be written: S(V 2) =(cid:20) V 2 1 − V 2 V 2 2 − V 2 (cid:21)t C −1(cid:20) V 2 1 − V 2 V 2 2 − V 2 (cid:21) = Y tC −1Y. It can be shown that the minimum is reached for: V 2 = (X tC −1X)−1X tC −1(cid:20) V 2 2 (cid:21) . 1 V 2 with X =(cid:20) 1 1(cid:21) , the uncertainty on V 2 being: V 2 = (X tC −1X)−1X tC −1. σ2 (18) (19) (20) (21) The above equations yield the final visibility estimate: V 2 = V 2 1 (σ2 2 − ρ12σ1σ2) + V 2 1 − ρ12σ1σ2) 1 (σ2 2 − 2ρ12σ1σ2 σ2 1 σ2 and the associated error: 1 σ2 2 (1 − ρ2 12)σ2 σ2 V 2 = σ2 1 + σ2 2 − 2ρ12σ1σ2 (22) (23) If ρ12 = 1 then the two single-output squared visibilities V 2 1 and V 2 2 are fully correlated and the above expression does not apply. In this case V is equal to one of the two single-output visibilities with its associated error bar. The quality of the fit is expressed by the χ2: χ2 = (V 2 1 + σ2 σ2 2 )2 1 − V 2 2 − 2ρ12σ1σ2 (24) This parameter is important because it allows us to check the consistency of the instrument and of the method to measure the visibilities and the error bars. If all assump- tions are correct then the χ2 should be equal to 1 on aver- age. In the FLUOR software we use this number as a data quality parameter. Data with χ2 greater than 3 should be examined in detail and rejected for science programs requiring a very good quality of calibration as the proba- bility to get a value larger than 3 is only of 8.33%. 5.5. Correlations of multiple baseline interferometer simultaneous visibilities Coherence factors recorded simultaneously on different baselines with telescopes in common may also be corre- lated. This correlation should be taken into account and saved with the reduced data in the form of a correlation matrix. The correlations may be as high as the correla- tions between the two channels of a coaxial interferometer as all calibrators are common to all baselines. The method used in Sect. 5.2 should be applied. A correlation matrix for the µ2 should be computed first. The final correlation factors for the final visibility estimates are then computed with a Monte-Carlo method. 6. Correlations between non-simultaneous visibilities Visibilities obtained on different baselines or on different days are usually considered independent. In the last para- graph, we focused on the possible correlations of visibili- ties recorded simultaneously on baselines with telescopes in common. In this section, we will consider the correla- tion due to common uncertainties in the calibration pro- cess for independent baselines or for visibilities measured at different times. The calibration of the transfer function may have required us to use the same calibrators hence the same diameter estimates. The errors on the visibilities are therefore not independent. It is the purpose of this para- graph to establish a method to compute this correlation and, more important, to be able to trace it to compute it a posteriori long after taking the data at the telescopes. Let S1 and S2 be two spatial frequencies at which squared visibilities V 2(S1) and V 2(S2) have been measured. The visibility estimates of Eqs. (10) and (22) can take the form: ( V 2(S1) = α1V a2(S1) + β1V b2 V 2(S2) = α2V c2(S2) + β2V d2 (S1) (S2) (25) where the calibrators are A, B, C and D. To save room, the calibrator visibilities are replaced by the capital letters. It can be shown that the correlation factor between the two squared visibilities is: ρ(V 2(S1), V 2(S2)) = × × × (26) α1α2pV ar(A)V ar(B)ρ(A, C) pV ar(V 2(S1))V ar(V 2(S2)) α1β2pV ar(A)V ar(D)ρ(A, D) pV ar(V 2(S1))V ar(V 2(S2)) β1α2pV ar(B)V ar(C)ρ(B, C) pV ar(V 2(S1))V ar(V 2(S2)) β1β2pV ar(B)V ar(D)ρ(B, D) pV ar(V 2(S1))V ar(V 2(S2)) When all calibrators are different, all correlations are zero. The correlation is maximum when a single calibrator has systematically been used. Measurement noise is present at the denominator only and the correlation is of course all the larger as the measurement noise is smaller. The correlation between two expected squared visibilities at two different baselines is not easy to evaluate analyti- cally. Besides, it may depend upon the model of the cali- brator. A computation can be performed which shows that the correlation is indeed equal to 1 with an excellent ac- curacy as long as no baseline is equal to 0. This can also 8 G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers be shown by expanding the visibility function. Thus, the expected visibilities derived from a uniform disk model of a same calibrator at two different baselines are fully cor- related to the first order. This holds as long as the second derivative of the model is small (which in the case of the uniform disk model is true except close to the zeros of the model) and as a condition, none of the baselines is very close to zero. In practice, the error on the diameter being usually small (less than 5%), the first order approxima- tion is valid and the two expected squared visibilities can therefore be considered fully correlated. This is true down to very short baselines as for example for a diameter of 10 ± 0.5 mas the correlation starts to decrease for a base- line below 5 cm. For practical use, Eq. (26) can be simplified as the corre- lations between expected visibilities are either 0 or 1 when the calibrators are respectively different or alike. The only requirement to compute this correlation is therefore that the variances of the expected visibilities and the coeffi- cients α and β be saved with the reduced data. These cor- relations will have to be computed to model fit the data. The generalization of the Levenberg-Marquardt method with correlated data is given in the Appendix at the end of the paper. 7. Validation of the method Examples of data reduction results and calibrations are presented. The quantities introduced in the previous sec- tions are discussed in practical situations and general com- ments on observing strategies are expressed. 7.1. Squared coherence factors statistics I have plotted in Fig. 1 three examples of µ2 distributions. In the case of V636 Her, the fringe speed puts the fringe frequency far above the turbulence piston spectrum. The piston is almost frozen during each scan and the amount of correlated noise is small. In the case of 71 UMa, the fringe speed is lower and the measurements are more sensitive to piston hence the higher correlated noise. δ Sge is an intermediate case. In all three examples, the distributions of µ2 are compatible with Gaussian distributions hence validating the basic assumption on the statistics of the µ2. An important fact is that the amount of correlated noise is not negligible and must be taken into account. However, a test on distributions is performed to detect deviations from Gaussian statistics. Deviations are not common and are always due to instrumental problems. In such cases, depending on the required level of data quality, data may be eliminated. 7.2. Examples of transfer functions Figure 2 presents two examples of squared co-transfer functions. Full circles are measurements on calibrators whereas open circles are interpolations for science targets. It is visible that the co-transfer function is not always sta- ble and may experience variations. In some cases like on May 15, 2000 at 8:07, an error of calibration may have happened as the co-transfer functions jump by a few per- cent. Yet, in most cases, the transfer functions variations are slow on time scales of a few hours and variations can be well approximated to the first order. Data collected on May 22, 2000 show that this is still the case when the cal- ibrator diameters are known with a very good precision. 7.3. Discussion of model fitting and examples 7.3.1. Amount of correlation Before presenting examples let us summarize the different levels of correlations we have encountered so far: 1. correlation of coherence factors (coaxial beamcombin- ers) 2. correlation of interferometric channels (coaxial beam- combiners) 3. correlation of simultaneous baselines 4. correlation of non-simultaneous baselines The first level (r) was addressed in Sect. 7.1. The amount of correlation between interferometric channels (ρ12) for a coaxial beamcombiner like FLUOR varies from a few percent for faint sources calibrated by very well-known calibrators to almost 100% for bright sources calibrated by sources whose diameters are known with an accuracy of a few percent. The two channels are therefore not fully independent in this case and it is important to check the χ2 defined by Eq. (24). A large χ2 may indicate that ei- ther the assumptions on Gaussian statistics were wrong for these particular data or that the transfer function varia- tion is not well measured. In either case, data should be examined in detail to decide whether the visibility value can be used or not. A blind method is to reject visibilities with a χ2 above a certain level that can be of 3 for difficult programs or relaxed to a larger value for easier programs. It is important to note that if the transfer function has varied accordingly in both channels at the time the sci- ence target was observed by an amount larger than the error bars then this χ2 test will fail to detect it. It can only be detected if the variations are opposite in the two channels. This is certainly a weakness. It will be interesting to assess the level of correlations of visibilities measured with multiple beam interferometers. It can be anticipated that it will not be negligible and will be of the same level as ρ12. The correlation between visibilities recorded separately is illustrated in Fig. 3. The data have been reduced in two different ways. Data plotted with open circles and fitted by a dashed-line uniform disk model are reduced without taking correlations into account. Data plotted with full circles and fitted by the continuous line were reduced with the method of this paper. In the first case, the fit is of very good quality with a χ2 smaller than 1. Yet, all visibilities have been calibrated with the same importance of G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers 9 source, hence a strong correlation between visibility values as the 3 × 3 correlation matrix shows: (27) C =  1 0.96 0.96 0.97 1 0.96 0.96 0.97 1   It is to be noticed that the ρ12 correlation factor is larger than 90% for all three visibilities, a large fraction of this correlation being due to the common calibrator. If corre- lations are ignored then noise is considered independent from one visibility to the other and this is why the first χ2 is smaller, as a large global noise is now interpreted as a large fluctuating noise from one visibility to the other. On the contrary, when correlations are used, a tiny fraction of noise (4% at most) can be considered a fluctuation giv- ing degrees of freedom for the adjustement of the model. This is equivalent to reducing error bars on visibilities by 96% in the fit. The common noise due to the uncertainty on the calibrator is then a simple common bias on the visibilities but does not contribute to the noise in the fit, hence the much larger χ2. In the zoomed part of this same figure, one can see that the fit now conforms to only one of the visibility data as the correlation matrix is close to being non-invertible (see the Appendix for the use of the correlation matrix in the fitting process). In more physical terms, the correlations being very close to one, all data are equivalent and the fit can be derived from one of the visi- bility data. If all data points were compatible despite the large correlations then the best fit curve would go through the error bars. It is not the case here and this is why the χ2 is large. 7.3.2. Examples of visibility accuracies Other examples of model fitting are presented in Fig. 4. The BK Vir data were calibrated with the same calibrator (the correlation matrix is similar to the matrix above) as SW Vir but the visibilities are very consistent with each others. Data for the three other sources are either totally independent or slightly correlated. Only the first lobe data were used for the fit of G Her. These four examples show very good fits and consistency of data. In particular, this sets the best absolute accuracy of the calibration of visi- bilities with FLUOR to 0.004 (equivalent to an accuracy of 0.004 on V 2 with V = 0.5 as σV 2 = 2σV .V to the first order). 7.3.3. Calibration strategies It is important to adapt the strategy of calibration to the type of astrophysical studies addressed with optical inter- ferometers. For most studies where visibility accuracies of a few percent are acceptable, the repeated use of a single or of a few calibrators is possible. For difficult programs like exoplanet detection, a very high level of accuracy is required and the strategy needs to be well prepared. Two cases may arise depending on whether the required cal- ibration of visibilities is absolute or relative. If absolute accuracies better than 0.001 have to be obtained on visi- bilities then it is very likely that no calibrator can be used twice, unless the error on the expected visibility of this calibrator is less than the level of accuracy required. This would suppose that the visibility model of the calibrator be measured first. Another possibility is relative detection. As illustrated by the example of SW Vir, if the same cal- ibrator is systematically used, the fit is sensitive to very low levels as the correlated noise does not contribute to the value of the χ2. In this example, a departure from the uniform disk model or a calibration error may have been detected to a level much lower than the error bars. For very faint detail detection, this can work if the visibility curve of the calibrator is smooth and without wiggles of similar amplitude as the ones searched for on the science target. In any case, the observing strategy should be prepared in advance and should take the problem of data correlations into account. 8. Conclusion I have proposed in this paper a method to calibrate visibil- ity data obtained with single mode interferometers. The single mode character is required to make valid the as- sumption that the statistics of coherence factors data are Gaussian and stationary. It is possible to derive reliable er- ror bars if all correlations are considered in the derivation of all estimators. Correlations also need to be taken into account when fitting the data by models. The validity of the method has been demonstrated on real interferomet- ric data recorded with FLUOR. An important conclusion of this work is that the strategy of calibration has to be adapted for specific programs requiring high standards of calibration. References Chelli, A., 2000, AMBER report AMB-IGR-017 Coud´e du Foresto, V., Ridgway, S.T., Mariotti, J.-M., 1997, A&ASS, 121, 379 Mourard, D., Tallon-Bosc, I., Rigal, F., et al. 1994, A&A 288, 675 Pelat, D., 1992, Cours "Bruits et Signaux", ´Ecole Doctorale d'Ile de France, Astronomie-Astrophysique Perrin, G., 1996, PhD thesis, Universit´e Paris VII Perrin, G., Coud´e du Foresto, V. Ridgway, S.T., et al., 1998, A&A 331, 619 Perrin, G., 2002, A&A, in press Press, W.H., Flannery, B.P., Teukolsky, S.A., Vetterling, W.T., 1988, "Numerical Recipes in C", Cambridge University Press Segransan, D., Forveille, T., Millan-Gabet, R., Perrier, C., Traub, W.A., 1999, "Working on the Fringe: Optical and IR Interferometry from Ground and Space", ASP Conference Vol. 194, p. 290, S. Unwin and R. Stachnik Eds. 10 G. Perrin: The calibration of interferometric visibilities obtained with single-mode optical interferometers Appendix A: Practical implementation of model fitting with correlated visibilities ∂2S ∂θk∂θl Algorithms for model fitting are well known. One of the most commonly used is the Levenberg-Marquardt algo- rithms. An example of practical implementation is given in Press et al. (1988) in the case where data are not cor- related. Here, I give a generalization of this algorithm to the case of correlated data. A.1. Definitions ∂θk(cid:21)t = (cid:20) ∂M C −1 ∂M ∂θl k, l = 1, . . . , p −(cid:20) ∂2M ∂θk∂θl(cid:21) C −1(cid:2)V 2 − M(cid:3) In case S is equal to its second order expansion then θ can be guessed directly through: θ = θ −(cid:20) ∂2S ∂θ2(cid:21) −1(cid:20) ∂S ∂θ(cid:21) (A.7) If the expansion is a poor approximation of S then a steep- est descent method has to be applied with: Let (V 2 i )i=1,...,N be a series of squared visibility measure- ments of an astronomical source. Let C be the matrix of variances-covariances of these measurements: θnext = θ − constant ×(cid:20) ∂S ∂θ(cid:21) (A.8) This the basis of the Levenberg-Marquardt method. It is classical to use the following notations: (A.1) βk = − ∂S ∂θk , k = 1, . . . , p ∂θk(cid:21)t αkl = (cid:20) ∂M C −1 ∂M ∂θl , k, l = 1, . . . , p where ∂M ∂θk is the kth column of the matrix: ∂M ∂θ =(cid:20) ∂M (xi, θ) ∂θk (cid:21)i,k (A.9) (A.10) The matrix equivalent of the above definition is therefore: ∂θ (cid:21)t β = (cid:20) ∂M ∂θ (cid:21)t α = (cid:20) ∂M C −1(cid:2)V 2 − M(cid:3) C −1 ∂M ∂θ (A.11) (A.2) At optimum the variance-covariance matrix of the param- eters is: (A.3) V ar(θ) = [α]−1 (A.12) σ2 1 ρ1,2σ1σ2 ... ρ1,N σ1σN C =  ρ1,2σ1σ2 · · · ρ1,N σ1σN σ2 2 ... · · · · · · . . . · · · ... ... σ2 N   Let (Mi(θ))i=1,...,N be the model to fit the data with θ a vector of p parameters. It is assumed that the noise on the squared visibilities is Gaussian. A.2. Method The best estimates of the parameters in the sense of the highest likelihood are obtained by maximizing the likeli- hood function of the Y = [V 2 − M (θ)] vector which leads to minimizing the following functional: S(θ) = Y tC −1Y 1 2 The optimum θ is obtained when: (cid:26) ∂S ∂θ = 0 ∂ 2S ∂θ2 > 0 With these definitions, 2S(θ) is a χ2 with N − p degrees of freedom. The increment in parameter space is therefore the solution of the linear system: αklδθl = βk p Xl=1 (A.13) when the quadratic form is a good approximation to S otherwise the increment is given by: δθl = constant × βl (A.14) The reader is referred to Press et al. (1988) for the imple- mentation of this modified algorithm. Besides, the covariance matrix of the estimated parame- ters is asymptotically (N → +∞) equal to: −1 (A.4) V ar(θ) =(cid:20) ∂2S ∂θ2(cid:12)(cid:12)(cid:12)(cid:12)θ= θ(cid:21) The search for the optimum can be performed with a gen- eralized Levenberg-Marquardt algorithm. A.3. Generalized Levenberg-Marquardt algorithm Close to minimum, S can be expanded to the second order in θ: ∂θ(cid:21)t S(θ) = S(θ)+(cid:20) ∂S 1 hθ − θi+ 2hθ − θit(cid:20) ∂2S ∂θ2(cid:21)hθ − θi (A.5) The first and second derivative of S are respectively a vec- tor and a matrix whose elements are: ∂S ∂θk ∂θk(cid:21)t = −(cid:20) ∂M C −1(cid:2)V 2 − M(cid:3) , k = 1, . . . , p (A.6)
astro-ph/0310193
1
0310
2003-10-07T20:00:33
Dark Matter in Dwarf Galaxies: Latest Density Profile Results
[ "astro-ph" ]
We present high-resolution two-dimensional velocity fields in Halpha and CO of the nearby dwarf galaxy NGC 2976. Our observations were made at both higher spatial resolution (~75 pc) and higher velocity resolution (13 km/s in Halpha and 2 km/s in CO) than most previous studies. We show that NGC 2976 has a very shallow dark matter density profile, with rho(r) lying between rho ~ r^-0.3 and rho ~ r^0. We carefully test the effects of systematic uncertainties on our results, and demonstrate that well-resolved, two-dimensional velocity data can eliminate many of the systematic problems that beset longslit observations. We also present a preliminary analysis of the velocity field of NGC 5963, which appears to have a nearly NFW density profile.
astro-ph
astro-ph
Satellites and Tidal Streams ASP Conference Series, Vol. **VOLUME***, 2003 F. Prada, D. Mart´ınez-Delgado, and T. Mahoney, eds. Dark Matter in Dwarf Galaxies: Latest Density Profile Results Joshua D. Simon, Alberto D. Bolatto, Adam Leroy, and Leo Blitz Department of Astronomy, University of California at Berkeley, 601 Campbell Hall, CA 94720 Abstract. We present high-resolution two-dimensional velocity fields in Hα and CO of the nearby dwarf galaxy NGC 2976. Our observations were made at both higher spatial resolution (∼ 75 pc) and higher velocity resolution (13 km s−1 in Hα and 2 km s−1 in CO) than most previous studies. We show that NGC 2976 has a very shallow dark matter density profile, with ρ(r) lying between ρ ∝ r−0.3 and ρ ∝ r0. We carefully test the effects of systematic uncertainties on our results, and demonstrate that well-resolved, two-dimensional velocity data can eliminate many of the systematic problems that beset longslit observations. We also present a preliminary analysis of the velocity field of NGC 5963, which appears to have a nearly NFW density profile. 1. Introduction It is well-known by now that there is a substantial disagreement between the observed dark matter density profiles of many dwarf and low-surface bright- ness galaxies and the density profiles predicted by numerical Cold Dark Matter (CDM) simulations (e.g., Flores & Primack 1994; Burkert 1995; Navarro, Frenk, & White 1996, hereafter NFW; Moore et al. 1999). The significance of this disagreement, though, remains controversial. A number of authors attribute the problem to failures of the simulations, or of the CDM model itself (de Blok et al. 2001a; de Blok, McGaugh, & Rubin 2001b; Borriello & Salucci 2001; de Blok, Bosma, & McGaugh 2003), while others argue that systematic uncertainties in the observations make such conclusions premature (van den Bosch et al. 2000; van den Bosch & Swaters 2001; Swaters et al. 2003, hereafter SMVB). We address this controversy with a new study that combines a number of techniques to overcome the systematics in the observations. We are acquiring very high-quality data on a limited sample of galaxies to investigate the impor- tance of systematic effects in detail. The results of this study should make clear whether systematic problems in the data are at fault, or whether there actually is a fundamental conflict between the theory and the observations. Our program includes (1) two-dimensional velocity fields obtained at opti- cal (Hα), millimeter (CO), and centimeter (H i) wavelengths, (2) high angular resolution (∼ 5′′), (3) high spectral resolution (∼< 10 km s−1), (4) multicolor optical and near-infrared photometry, and (5) nearby dwarf galaxies as targets. 1 2 Simon et al. The combination of these features greatly reduces our vulnerability to systematic uncertainties (see Simon et al. 2003). In two previous papers, we reported on rotation curve studies of the dwarf spiral galaxies NGC 4605 and NGC 2976 (Bolatto et al. 2002; Simon et al. 2003). In this paper we highlight some of those results, focusing on tests for systematic uncertainties in the NGC 2976 data set, and we also present preliminary results on the third galaxy in our study, NGC 5963. 2. The Dark Matter Halo of NGC 2976 NGC 2976 is a regular Sc dwarf galaxy located in the M81 group at a distance of 3.45 Mpc. The galaxy has absolute magnitudes of MB = −17.0 and MK = −20.2, and a total mass of 3.5 × 109M⊙; it is somewhat less luminous and less massive than the Large Magellanic Cloud. In optical and near-infrared images it is clear that NGC 2976 is a bulgeless, unbarred, pure disk system, which makes it an ideal galaxy for mass modeling. 2.1. Observations Our observations of NGC 2976 include a two-dimensional Hα velocity field (obtained with a multi-fiber spectrograph on the WIYN telescope), a two- dimensional CO velocity field (obtained with the BIMA interferometer), multi- color optical photometry, and near-infrared 2MASS imaging. The velocity fields are both shown in Figure 1. For details of the data reduction and analysis, see Simon et al. (2003). 2.2. Rotation Curve We used various tilted-ring modeling algorithms to convert the observed veloc- ity field into a rotation curve (see Simon et al. 2003). The rotation curve of NGC 2976 is well-described by a single power law from the center of the galaxy out to a radius of almost 2 kpc, as displayed in Figure 2a. The rotation curve only begins to deviate systematically from power law behavior at r ≈ 110′′ (1.84 kpc). The total (baryonic plus dark matter) density profile corresponding to the rotation curve is ρTOT = 1.6(r/1 pc)−0.27±0.09M⊙ pc−3. Because the baryons are almost certainly centrally concentrated, this density profile represents the cuspiest possible shape for the dark matter halo. 2.3. Dark Matter Density Profile To explore the full range of possible dark matter density profile slopes, we must account for the rotational velocities contributed by the baryonic components of the galaxy. Because our images of NGC 2976 do not reveal a bulge or a bar, and its nucleus is dynamically unimportant, the only relevant reservoirs of baryons to consider are the stellar and gaseous disks. We calculated the rotation curves due to the stars and gas directly from the observed surface density profiles, assuming an infinitely thin disk. The only free parameter in these calculations is the K-band mass-to-light ratio of the stellar population (M∗/LK ). Under the assumption that the density profile can be described with a power law, ρDM ∝ r−αDM, we found that if M∗/LK > 0.19M⊙/L⊙K, αDM < 0 and the Latest Density Profile Results 3 100 a) b) 80 60 40 20 0 ) " ( t e s f f O . c e D −20 −40 −60 −80 −100 50 0 −50 R.A. Offset (") 50 0 −50 R.A. Offset (") VHEL (km s−1) 80 60 40 20 0 −20 −40 −60 −80 Figure 1. (a) Hα velocity field of NGC 2976 from WIYN observa- tions. The contours represent integrated Hα intensity. (b) CO velocity field from BIMA observations. The contours represent integrated CO intensity. The angular resolution of each dataset is shown in the upper left corners. density of the dark matter halo is increasing with radius. Such a dark matter con- figuration is probably unphysical, so we consider 0.19 M⊙/L⊙K to be a firm up- per limit to the stellar disk mass-to-light ratio, with the corresponding lower limit to αDM of 0. Because of the extremely low value of the maximal disk mass-to- light ratio, the galaxy must contain an essentially maximal disk. The dark mat- ter density profile for the maximal disk is ρDM = 0.1(r/1 pc)−0.01±0.13M⊙ pc−3. This represents the most likely shape for the dark matter halo. Since the slope of the total density profile of the galaxy represents the absolute upper limit for the slope of the dark matter density profile, we conclude that the dark matter density profile is bracketed by ρDM ∝ r−0.27±0.09 and ρDM ∝ r0. 3. Systematics Several authors have recently discussed the systematic uncertainties that can significantly alter observed rotation curves. SMVB used simulated observations to argue that systematic effects alone could account for the difference between the predicted and observed density profile slopes. de Blok et al. (2003) car- ried out similar simulations, but concluded that the systematic effects were not strong enough to cause cuspy density profiles to appear to have cores. We have shown that, for the specific case of NGC 2976, the known systematic problems do not distort the rotation curve (Simon et al. 2003; this work). The worst systematics can probably be minimized or avoided in general by using two- 4 Simon et al. 90 80 70 60 50 40 30 20 10 0 0 −10 −20 500 a) 0 20 ) 1 − s m k ( n o i t t a o r V ) 1 − s m k ( l a u d s e r i V Galactocentric Radius (pc) 1000 1500 60 40 100 Galactocentric Radius (") 80 2000 500 b) 120 140 0 20 Galactocentric Radius (pc) 1000 1500 60 40 100 Galactocentric Radius (") 80 2000 120 140 Figure 2. (a) Minimum disk rotation curve of NGC 2976. The ob- served rotation velocities are plotted as black circles and the error bars are combined statistical and systematic uncertainties. The dashed and dotted curves represent the rotation velocities due to H i and H2, re- spectively. A power law fit to the rotation curve is shown by the solid black curve. The corresponding density profile is ρ ∝ r−0.27. Residuals from the fit are displayed in the lower panel, and 1σ and 2σ depar- tures from the fit are represented by the shaded regions. (b) Maximum disk rotation curve of NGC 2976. The stellar disk (thick gray curve) is scaled up as high as the observed rotation velocities (gray circles) allow. The stellar disk shown here has M∗/LK = 0.19 M⊙/L⊙K . Af- ter subtracting the rotation velocities due to the stars and the atomic and molecular gas from the observed rotation curve, the dark mat- ter rotation velocities are displayed as black circles. The two missing data points near the center of the galaxy had vrot < v∗,rot, yielding imaginary vhalo. The solid black line is a power law fit to the halo ve- locities (for 14′′ < r < 105′′) which corresponds to a density profile of ρDM ∝ r−0.01. The halo residuals after the power law fit are displayed in the bottom panel. dimensional velocity fields and by making velocity measurements at very high precision (Blais-Ouellette et al. 1999; van den Bosch & Swaters 2001; Bolatto et al. 2002; SMVB), but additional observations confirming this expectation for more galaxies are needed. 3.1. Rotation Curve Fitting Systematics Because of the high precision of our velocity measurements, the statistical un- certainties on both the rotation curve and the radial velocity curve are negligible (less than 1 km s−1 everywhere). Therefore, the limiting factors on the accuracy of the rotation curve are the systematic uncertainties associated with our fit. Latest Density Profile Results 5 Uncertainties in Geometric Parameters Before computing the uncertainties on the rotation velocities themselves, we must first determine the uncertainties on each of the parameters that are used to calculate the rotation velocities: the center, position angle (PA), inclination, and systemic velocity. We used a bootstrap resampling technique to calculate the value and uncertainty of these parameters. We constructed 200 bootstrap samples of the velocity field and ran our tilted-ring modeling routine (ringfit) on each of them to determine the best-fit kinematic center. We then defined the center of the galaxy to be the median of these 200 measurements, and the systematic uncertainty on the center position to be the dispersion of the measurements. We found that the location of the kinematic center of NGC 2976 is consistent with the optical nucleus, with an uncertainty of 2′′ in both α and δ. We used the same bootstrap method to measure the kinematic PA of the galaxy and its uncertainty, finding that the kinematic PA matches the photometric one and has an uncertainty of 5◦. It is not possible to determine a kinematic inclination angle for NGC 2976 because the rotation curve is too close to solid-body. However, the photometric inclination angle is well determined, and we estimate that the uncertainty is 3◦. The uncertainty in the systemic velocity was calculated in a different way. The systemic velocity of each ring was left free to vary during our tilted-ring fitting, resulting in 23 independent measurements. The overall uncertainty from these measurements is 1.8 km s−1. Since vsys is always a free parameter, this uncertainty does not directly factor in to the uncertainty on the rotation curve. Uncertainties in Rotation Velocities and Radial Velocities Using the measured uncertainties in the center position, PA, and inclination angle, we calculated the resulting uncertainties on the rotation velocities and the radial velocities with a Monte Carlo study. We generated 1000 random centers, PAs, and inclinations, assuming a Gaussian distribution for each of the parameters, and ran ringfit with each set of parameters. The standard deviation of the 1000 rotation veloc- ities in each ring was defined to be the systematic error of that rotation velocity, and the systematic errors in the radial velocities and systemic velocities were calculated in the same way. The systematic errors on the rotation curve range from 2.1 km s−1 to 5.5 km s−1. The sum in quadrature of these errors and the statistical errors translates to a total uncertainty on the density profile slope α of 0.09. We conclude that, even in the minimum disk case, an NFW or steeper density profile in NGC 2976 is strongly ruled out. 3.2. Other Systematic Effects In some studies, beam-smearing is one of the dominant systematic effects. The velocity field of NGC 2976, though, is extremely well-resolved -- our spatial resolution is less than 100 pc in both CO and Hα. The measured velocity gradient is small (∼ 50 km s−1 kpc−1) and the constant-density core that we detected is ∼ 50 resolution elements across, indicating that beam-smearing is not responsible for this result. We also calculated the asymmetric drift correction to the rotation curve. Its effect is to increase the maximum disk mass-to-light ratio, and to make the rotation curve slightly more linear, but the density profile does not change significantly. 6 Simon et al. 3.3. Comparing Velocities Derived From Different Tracers Some recent studies in the literature have shown that, beam smearing questions aside, there do not appear to be systematic offsets between H i and Hα rotation velocities (e.g., McGaugh, Rubin, & de Blok 2001; Marchesini et al. 2002). With a handful of exceptions, though, these studies employed longslit Hα data, so the comparisons essentially took place only along the major axis. In addition, the spatial and velocity resolution of the H i and Hα data were often quite different. Simon et al. (2003) presented for the first time the data necessary for a two- dimensional comparison across a dwarf galaxy of the CO and Hα velocity fields. The angular resolution of the two datasets is similar (6′′ and 4′′, respectively), and although the CO velocity resolution is better by a factor of ∼ 6, the higher signal-to-noise at Hα enables us to measure the velocities with comparable pre- cision. We constructed a unique one-to-one mapping between the two velocity fields. The mean difference between vHα and vCO is 1 km s−1 and the rms is 5.3 km s−1, with the comparison being made at 173 independent points. Sim- ilar studies in the Milky Way found that the dispersion between the velocities of molecular clouds and the associated Hα-emitting gas was 4-6 km s−1 (Fich, Dahl, & Treffers 1990), so much of the scatter we observe between the two in NGC 2976 may be intrinsic to the process of H ii region formation rather than caused by observational uncertainties. This comparison strongly suggests that both the CO and the Hα data accu- rately reflect the gravitational potential of the galaxy, and neither is significantly compromised by systematic effects. Extinction, for example, is not distorting the Hα velocity field. If this result can be shown to hold for a few additional galaxies, then it will no longer be necessary to observe more than one velocity field tracer per galaxy. One further test that should be carried out is to compare the velocity field from a gaseous emission-line tracer to that from a stellar absorption-line tracer to assess the agreement between the stellar and gas kinematics. Such stud- ies have been done before (e.g., Mulder 1995; Bottema 1999), but usually only with long-slit observations. Obtaining two-dimensional absorption-line velocity fields is currently observationally challenging but feasible. 3.4. Noncircular Motions Perhaps the most surprising result from Simon et al. (2003) was the finding that NGC 2976 contains strong radial motions. In optical and near-infrared images, the galaxy is very regular, with no hint of a bar or any other deviation from axisymmetry that might affect the kinematics. The finding that the velocity field is clearly different from a purely rotating disk was therefore unexpected. Near the center of the galaxy, the direction of the maximum velocity gradient is up to ∼ 40◦ away from the photometric major axis. The velocity data can be adequately described by a model in which the galaxy is undergoing circular rotation, but only if the kinematic PA declines monotonically from ∼ 6◦ to ∼ −37◦ over the central kiloparsec. As mentioned above, though, this position angle variation is inconsistent with the photometry. The alternative -- and more likely -- model is one in which NGC 2976 contains substantial radial flows. The origin of such motions, however, is unclear. There are several plausible mechanisms for creating radial motions. An intriguing possibility is that the radial motions could be a result of the dark Latest Density Profile Results 7 matter halo having a triaxial rather than spherical shape. CDM halos are ex- pected to be moderately triaxial (e.g., Dubinski & Carlberg 1991; Warren et al. 1992), and the velocity field of a galaxy embedded in a triaxial halo would exhibit noncircular motions. However, since the details of such a velocity field have not yet been investigated with simulations, we cannot compare our results to theoretical predictions. We plan to carry out simulations of the kinematics of a disk within a triaxial halo to see how the halo shape influences the gas kine- matics and whether the effect is consistent with the velocity field of NGC 2976. This could become a new technique for measuring the triaxiality of galaxy dark matter halos. 4. The Dark Matter Halo of NGC 4605 NGC 4605 is another Sc dwarf galaxy, located at a distance of 4.3 Mpc. Like NGC 2976, it appears to be a pure disk system. We used a long-slit Hα spec- trum and a two-dimensional CO velocity field to measure its rotation curve (Bolatto et al. 2002). We found that the galaxy probably contains a maximal disk, and that its dark matter density profile goes as ρ ∝ r−0.65 (see Fig. 3), intermediate between a core and a cusp. The quality of our original observations of NGC 4605 precluded an investigation of the systematics, but we are in the process of analyzing new observations to study their impact. 5. The Dark Matter Halo of NGC 5963 NGC 5963 is a more massive (∼ 1010M⊙) and more distant (∼ 10 Mpc) galaxy than NGC 2976 and NGC 4605. Its surface brightness profiles are more com- plicated than those galaxies as well, showing bright emission at the center, sur- rounded by a transition region in which the surface brightness falls off extremely rapidly (scale length ≈ 300 pc), and then a very low surface brightness expo- nential disk extending from 1.5 kpc out to beyond 3 kpc. The interpretation of this behavior near the center of the galaxy is not straightforward; the galaxy is definitely not strongly barred, but we cannot rule out a bulge component. 5.1. Observations We obtained a two-dimensional Hα velocity field of NGC 5963 with the WIYN telescope, using the same technique as for NGC 2976. The spatial resolution of these data is 190 pc and the velocity resolution is 13 km s−1. We detected Hα emission out to a radius of ∼ 3 kpc. We also obtained a two-dimensional CO velocity field from BIMA. The CO emission in this galaxy is limited to the central kiloparsec, but adding the CO data improves our ability to constrain the innermost part of the density profile. 5.2. Rotation Curve and Stellar Disk The combined CO and Hα rotation curve of NGC 5963 is displayed in Figure 4. It is immediately clear that the density profile of NGC 5963 has a very different structure from those of NGC 2976 and NGC 4605. Instead of a mostly linear 8 Simon et al. Galactocentric Radius (pc) 1000 1500 500 100 a) 2000 500 b) Galactocentric Radius (pc) 1000 1500 2000 ) 1 − s m k ( n o i t t a o r V 80 60 40 20 0 0 ) 1 − s m k ( l a u d s e r i V −10 −20 0 10 20 30 40 50 60 70 80 90 100 110 0 10 20 30 40 50 60 70 80 90 100 110 Galactocentric Radius (") Galactocentric Radius (") Figure 3. (a) Minimum disk rotation curve of NGC 4605. The black curve is a ρ ∝ r−1.1 power law fit, which agrees quite closely with the rotation curve beyond 25′′. At smaller radii, though, it signifi- cantly overestimates the rotation velocities. A separate inner power law (ρ ∝ r−0.4) is necessary to fit the entire rotation curve. The bot- tom panel shows the residuals from the fit, with 1σ and 2σ departures represented by the shaded regions. Statistical error bars are displayed in the residual plot. (b) Maximum disk rotation curve of NGC 4605. The light gray circles represent the observed rotation curve, and the black circles represent the rotation curve of the dark matter halo after the stellar disk contribution (thick gray curve) has been subtracted. The black curve is a ρ ∝ r−0.65 power law that provides a good fit to the dark matter rotation curve at all radii. increase with radius, the rotation curve of NGC 5963 has the classic spiral galaxy shape with a steep inner rise and a nearly flat outer portion. We perform isophotal fits to the I-band image of NGC 5963 to obtain the shape of its surface brightness profiles. We then calculate the rotation curve due to the stellar disk numerically, as described before. The resulting stellar disk rotation curve can be seen in Figure 4. The maximum disk limit for this galaxy is set by the innermost point of the rotation curve. We derive a maximum disk mass-to-light ratio of M∗/LI = 0.7M⊙/L⊙I . Even the maximum disk is unable to account for the observed rotation speed of the galaxy beyond a radius of ∼ 6′′. The calculated value of the maximum disk mass-to-light ratio is not unusually low, so a maximum disk is not required in NGC 5963. The disk can be less massive than shown in Figure 4, but since the galaxy is dark matter-dominated the structure of the dark halo does not depend strongly on the assumed stellar mass-to-light ratio. Latest Density Profile Results 9 500 a) 140 120 100 80 60 40 20 0 15 0 −15 0 10 Galactocentric Radius (pc) 1000 2000 1500 30 20 50 Galactocentric Radius (") 40 2500 3000 500 b) 60 0 10 Galactocentric Radius (pc) 1000 1500 2000 30 20 50 Galactocentric Radius (") 40 2500 3000 60 ) 1 − s m k ( n o i t t a o r V ) 1 − s m k ( l a u d s e r i V Figure 4. (a) Minimum disk rotation curve of NGC 5963. The black curve shows an NFW fit to the rotation curve with a concentration of 20. Residuals from the fit are plotted in the bottom panel. Un- like the previous two galaxies, a power law is not a very good fit and the NFW form fits much more closely. (b) Maximum disk rotation curve of NGC 5963. Even after removing a maximal stellar disk (thick gray curve; M∗/LI = 0.7M⊙/L⊙I ), the rotation curve due to the dark matter halo is well-fit by an NFW potential. 5.3. Dark Matter Density Profile Fitting a single power law to the rotation curve of NGC 5963 does not produce a very good fit. The fit is substantially improved by using separate inner and outer power laws, corresponding to an inner density profile of ρ ∝ r−1.1 and a steeper outer density profile of ρ ∝ r−1.4. An even better fit can be obtained by fitting an NFW rotation curve to the data; the reduced χ2 value of this fit is only 1.1. Although the concentration parameter of the halo is not well-constrained, the scale radius is about 3 kpc and the halo has v200 ≈ 90 km s−1. An NFW fit can accurately describe the rotation curve for any stellar mass-to-light ratio (see Fig. 4). This galaxy may be the first low-mass system found that contains a density profile that matches the predictions of the numerical simulations. 6. Conclusions We have used high-resolution two-dimensional velocity fields to study the dark matter density profiles of NGC 2976, NGC 4605, and NGC 5963. We showed that these three galaxies contain very different density profiles. For NGC 2976, we presented one of the most detailed velocity fields of a dwarf galaxy outside the Local Group. Even with these high-quality data, we found that NGC 2976 contains a constant-density core and an NFW halo is strongly ruled out. The 10 Simon et al. dark matter halo of NGC 4605 is intermediate between a constant-density core and a cusp, and NGC 5963 has a cuspy halo that is almost perfectly consistent with an NFW profile. We conclude that some galaxies do not contain central cusps, even when systematic uncertainties are accounted for as carefully as possible. We also point out that we find no evidence to support the prediction that all dark matter halos share a universal shape. If these results hold as our sample grows, it will demonstrate that the conflict between the observations and the simulations is not caused by systematic problems with the observations. Instead, more effort may be needed to investigate what could be missing from the simulations that would cause them to overestimate density profile slopes. Acknowledgments. This research was supported by NSF grant AST-9981308. References Blais-Ouellette, S., Carignan, C., Amram, P., & Cot´e, S. 1999, AJ, 118, 2123 Bolatto, A. D., Simon, J. D., Leroy, A., & Blitz, L. 2002, ApJ, 565, 238 Borriello, A., & Salucci, P. 2001, MNRAS, 323, 285 Bottema R. 1999, A&A, 348, 77 Burkert, A. 1995, ApJ, 447, L25 de Blok, W. J. G., Bosma, A., & McGaugh, S. 2003, MNRAS, 340, 657 de Blok, W. J. G., McGaugh, S. S., Bosma, A., & Rubin, V. C. 2001a, ApJ, 552, L23 de Blok, W. J. G., McGaugh, S. S., & Rubin, V. C. 2001b, AJ, 122, 2396 Dubinski, J., & Carlberg, R. G. 1991, ApJ, 378, 496 Fich, M., Dahl, G. P., & Treffers, R. R. 1990, AJ, 99, 622 Flores, R. A., & Primack, J. R. 1994, ApJ, 427, L1 Marchesini, D., D'Onghia, E., Chincarini, G., Firmani, C., Conconi, P., Moli- nari, E., & Zacchei, A. 2002, ApJ, 575, 801 McGaugh, S. S., Rubin, V. C., & de Blok, W. J. G. 2001, AJ, 122, 2381 Moore, B., Quinn, T., Governato, F., Stadel, J., & Lake, G. 1999, MNRAS, 310, 1147 Mulder, P. S. 1995, A&A, 303, 57 Navarro, J. F., Frenk, C. S., & White, S. D. M. 1996, ApJ, 462, 563 (NFW) Simon, J. D., Bolatto, A. D., Leroy, A., & Blitz, L. 2003, ApJ, in press (preprint: astro-ph 0307154) Swaters, R. A., Madore, B. F., van den Bosch, F. C., & Balcells, M. 2003, ApJ, 583, 732 (SMVB) van den Bosch, F. C., Robertson, B. E., Dalcanton, J. J., & de Blok, W. J. G. 2000, AJ, 119, 1579 van den Bosch, F. C., & Swaters, R. A. 2001, MNRAS, 325, 1017 Warren, M. S., Quinn, P. J., Salmon, J. K., & Zurek, W. H. 1992, ApJ, 399, 405
astro-ph/0604036
1
0604
2006-04-03T19:28:32
Analysis of 26 Barium Stars I. Abundances
[ "astro-ph" ]
We present a detailed analysis of 26 barium stars, including dwarf barium stars, providing their atmospheric parameters (Teff, log g, [Fe/H], vt) and elemental abundances. We aim at deriving gravities and luminosity classes of the sample stars, in particular to confirm the existence of dwarf barium stars. Accurate abundances of chemical elements were derived. Abundance ratios between nucleosynthetic processes, by using Eu and Ba as representatives of the r- and s-processes are presented. High-resolution spectra with the FEROS spectrograph at the ESO-1.5m Telescope, and photometric data with Fotrap at the Zeiss telescope at the LNA were obtained. The atmospheric parameters were derived in an iterative way, with temperatures obtained from colour-temperature calibrations. The abundances were derived using spectrum synthesis for Li, Na, Al, alpha-, iron peak, s- and r-elements atomic lines, and C and N molecular lines. Atmospheric parameters in the range 4300 < Teff < 6500, -1.2 < [Fe/H] < 0.0 and 1.4 < log g < 4.6 were derived, confirming that our sample contains giants, subgiants and dwarfs. The abundance results obtained for Li, Al, Na, alpha- and iron peak elements for the sample stars show that they are compatible with the values found in the literature for normal disk stars in the same range of metallicities. Enhancements of C, N and heavy elements relative to Fe, that characterise barium stars, were derived and showed that [X/Ba] vs. [Ba/H] and [X/Ba] vs. [Fe/H] present different behaviour as compared to [X/Eu] vs. [Eu/H] and [X/Eu] vs. [Fe/H], reflecting the different nucleosynthetic sites for the s- and r-processes.
astro-ph
astro-ph
Astronomy&Astrophysicsmanuscript no. bario1 April 26, 2018 c(cid:13) ESO 2018 Analysis of 26 Barium Stars ⋆,⋆⋆ I. Abundances D.M. Allen⋆⋆⋆ and B. Barbuy 6 0 0 2 r p A 3 1 v 6 3 0 4 0 6 0 / h p - o r t s a : v i X r a Instituto de Astronomia, Geof´ısica e Ciencias Atmosf´ericas, Universidade de Sao Paulo, Rua do Matao 1226, 05508-900 Sao Paulo, Brazil, e-mail: [email protected], [email protected] Received, 2006; accepted, 2006 ABSTRACT Context. We present a detailed analysis of 26 barium stars, including dwarf barium stars, providing their atmospheric parameters (Teff, log g, [Fe/H], vt) and elemental abundances. Aims. We aim at deriving gravities and luminosity classes of the sample stars, in particular to confirm the existence of dwarf barium stars. Accurate abundances of chemical elements were derived. Abundance ratios between nucleosynthetic processes, by using Eu and Ba as representatives of the r- and s-processes are presented. Methods. High-resolution spectra with the FEROS spectrograph at the ESO-1.5m Telescope, and photometric data with Fotrap at the Zeiss telescope at the LNA were obtained. The atmospheric parameters were derived in an iterative way, with temperatures obtained from colour-temperature calibrations. The abundances were derived using spectrum synthesis for Li, Na, Al, α-, iron peak, s- and r-elements atomic lines, and C and N molecular lines. Results.Atmospheric parameters in the range 4300 < Te f f < 6500, -1.2 < [Fe/H] < 0.0 and 1.4 ≤ log g < 4.6 were derived, confirming that our sample contains giants, subgiants and dwarfs. The abundance results obtained for Li, Al, Na, α- and iron peak elements for the sample stars show that they are compatible with the values found in the literature for normal disk stars in the same range of metallicities. Enhancements of C, N and heavy elements relative to Fe, that characterise barium stars, were derived and showed that [X/Ba] vs. [Ba/H] and [X/Ba] vs. [Fe/H] present different behaviour as compared to [X/Eu] vs. [Eu/H] and [X/Eu] vs. [Fe/H], reflecting the different nucleosynthetic sites for the s- and r-processes. Key words. barium stars -- atmospheric parameters -- abundances 1. Introduction Barium stars were recognized as a distinct group of peculiar stars by Bidelman & Keenan (1951). Initially, the objects in- cluded in this group were only G and K giants which showed strong lines of s-process elements, particularly Ba II at λ4554 Å and Sr II at λ4077 Å, as well as enhanced CH, CN and C2 molecular bands. However, the discovery by Tomkin et al. (1989) that the dwarf star HR 107 showed chemical compo- sition similar to that of a mild barium giant, has pushed the Send offprint requests to: D.M. Allen ⋆ Based on spectroscopic observations collected at the European Southern Observatory (ESO), within the Observat´orio Nacional ON/ESO and ON/IAG agreements, under FAPESP project n◦ 1998/10138-8. Photometric observations collected at the Observat´orio do Pico dos Dias (LNA/MCT). ⋆⋆ Tables 5 and 15 are only available in elecronic at the CDS via anonymous ftp to cdsarc.u-strasbg.fr/Abstract.html ⋆⋆⋆ Present address: Observat´orio do Valongo/UFRJ, Ladeira do Pedro Antonio 43, 20080-090 Rio de Janeiro, RJ, Brazil search for less evolved barium stars (G´omez et al. 1997). Some studies have proposed that these stars could be ancestors of the barium giants (North et al. 1994; Barbuy et al. 1992). McClure et al. (1980) revealed that most barium stars, maybe all of them, show variations in radial-velocity sug- gesting the presence of companions. This has been con- firmed by McClure (1983, 1984) and Udry et al. (1998a,b). Bohm-Vitense (1980) and Bohm-Vitense & Johnson (1985) observed an ultraviolet excess in the barium stars ζ CAP and ξ Ceti, which could be explained by white dwarf companions. The binarity hypothesis for all barium stars has provided an interesting explanation for their peculiarity. In this context, the more massive of them evolves through the thermal pulse - asymptotic giant branch (TP-AGB) phase, when s-process occurs, and afterwards the third dredge-up brings to the sur- face carbon and s-process elements. The enriched material is then transferred to the companion, which becomes a barium star. Bond et al. (2003) observed a planetary nebula (PN) in Cassiopeia, with a late type star, showing overabundance in 2 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars carbon and s-process elements typical of a barium star. This discovery confirmed the hypothesis about their origin. In the PNs Abell 35 (Jacoby 1981; Th´evenin & Jasniewicz 1997) and LoTr5 (Jasniewicz et al. 1996; Th´evenin & Jasniewicz 1997) barium and s-elements rich stars were also observed. There is a central hot star in each of these nebulae detected with IUE (International Ultraviolet Explorer). Jeffries & Smalley (1996) observed the binary system 2REJ0357+283 consisting of a white dwarf and an s-element rich K dwarf star, where the high rotation can be attributed to the mass transfer from the white dwarf progenitor, during the AGB phase. However, several IUE spectra of barium stars analysed by Dominy & Lambert (1983) showed no UV excess, putting in check the hypothesis that all barium stars have a white dwarf companion. Bohm-Vitense et al. (2000) observed UV excess for most of their barium stars, but the estimated cooling time for some of the companion white dwarfs, was too long or larger than the evolution time of the barium star. If the binarity hy- pothesis is not confirmed for all stars of the group, another ex- planation for their origin would be needed. In Sect. 2 the observations are reported. In Sect. 3 the at- mospheric parameters are derived. In Sect.4 the abundances derivation is described. In Sect. 5 conclusions are drawn. 2. Observations The sample stars were selected from G´omez et al. (1997) and North et al. (1994) where the authors suggested that there were less evolved barium stars among the giants of their sample. The star HR 107 from Tomkin et al. (1989) was also included in the sample. Mennessier et al. (1997) identified HD 5424, HD 13551, HD 116869 and HD 123396 as halo stars. Optical spectra were obtained at the 1.52m telescope at ESO, La Silla, using the Fiber Fed Extended Range Optical Spectrograph (FEROS) (Kaufer et al. 2000), on February and October/2000, January and October/2001, January and July/2002. The total spectrum coverage is 356-920 nm with a resolving power of 48,000. Two fibers, with entrance aper- ture of 2.7 arcsec, recorded simultaneously star light and sky background. The detector is a back-illuminated CCD with 2948×4096 pixels of 15 µm size. Reductions were carried out through a special pipeline package for reductions (DRS) of FEROS data, in MIDAS environment. The data reduction pro- ceeded with subtraction of bias and scattered light in the CCD, orders extraction, flatfielding, and wavelength calibration with a ThAr calibration frame. Radial velocities were taken into ac- count by using IRAF tasks RVIDLINE and DOPCOR. The spectra were cut in parts of 100 Å each using the SCOPY task, and the normalization was carried out with the CONTINUUM task. The photometric observations were obtained using the FOTRAP (Fotometro R´apido) at the ZEISS 60cm telescope at LNA (Laborat´orio Nacional de Astrof´ısica) in June, August and September/2001 and May and July/2002. Data reductions were done using the appropriate code available at LNA. The colours obtained were (B-V), (V-R), (R-I) and (V-I). The log of observations is presented in Table 1. Table 1. Log of spectroscopic (sp) and photometric (pho) observations. The S/N ratio was measured in the of λ5000 Å region. Radial velocity vr (km/s) is shown in column 6. References used to build this sample: 1 - G´omez et al. (1997); 2 - North et al. (1994); 3 - Tomkin et al. (1989). star HD749 HR107 HD5424 HD8270 HD12392 HD13551 HD22589 HD27271 HD48565 HD76225 HD87080 HD89948 HD92545 HD106191 HD107574 HD116869 HD123396 HD123585 HD147609 HD150862 HD188985 HD210709 HD210910 HD222349 BD+185215 HD223938 datepho datesp Exp. (s) S/N vr ref ... 31/08/2001 17/01/2001 16/07/2002 05/10/2001 01/09/2001 17/01/2001 01/09/2001 14/02/2000 31/08/2001 17/01/2001 01/09/2001 14/02/2000 01/09/2001 14/02/2000 01/09/2001 14/02/2000 23/01/2002 18/05/2002 23/01/2002 18/05/2002 14/02/2000 18/05/2002 14/02/2000 18/05/2002 23/01/2002 19/05/2002 03/07/2002 18/05/2002 23/01/2002 18/05/2002 14/02/2000 19/05/2002 15/02/2000 18/05/2002 04/07/2002 18/05/2002 04/07/2002 18/05/2002 04/07/2002 19/05/2002 04/10/2001 26/06/2001 18/10/2000 27/06/2001 19/10/2000 15/07/2002 02/10/2001 05/10/2001 15/07/2002 18/10/2000 ... 1800 200 15.47 1 9.76 3 600 200 3000 100 -1.56 1 1800 150 14.97 1 2700 120 -25.38 1 2700 100 34.79 1 2700 200 -28.90 1 1800 250 -16.65 1 600 250 -33.67 2 -1.34 2 1200 150 8.10 1 2700 120 6.01 1 1800 250 -0.16 2 1200 150 3600 100 0.68 2 0.78 2 1320 200 3600 150 -9.84 1 3600 150 26.64 1 7200 100 27.87 1,2 3600 120 -19.39 2 3600 150 50.38 2 1800 130 2700 100 27.01 1 2400 200 -9.86 1 2400 150 27.65 2 3600 100 -33.98 2 2700 250 -0.85 1 8.17 1,2 3. Atmospheric Parameters 3.1. Extinction Reddening values shown in Table 3 were derived accord- ing to Cardelli et al. (1989), using effective wavelengths by Bessell & Brett (1988) and Bessell (1979). The visual extinc- tion Av was considered null for stars nearer than 70 pc accord- ing to Vergely et al. (1998), and for the other stars, Av was de- terminated according to Chen et al. (1998). Distances were taken from Mennessier et al. (1997). Including Hipparcos parallaxes as input, these authors used a Maximum Likelihood method, which they considered to provide better results for the distances than the ones ob- tained directly from the parallaxes. The distances missing in Mennessier et al., were obtained directly from Hipparcos par- allaxes. In cases where no parallax values were available in the Hipparcos Catalogue, initial distances were estimated as- suming that these stars are subgiants of absolute magnitude Mv = 3.3 from G´omez et al. (1997), or dwarfs of Mv = 4.5. The V magnitudes in these cases were taken from SIMBAD available at the web address http://simbad.u-strasbg.fr/Simbad. Distances are shown in Table 2. D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 3 Table 2. Equatorial and galactic coordinates of the sample barium stars. The errors on last decimals are given in parenthesis. Distances DM are from Mennessier et al. (1997); distances DH were determined from Hipparcos parallaxes (DH =1/πH); Dg are the distances from spectroscopy for stars with Hipparcos parallax not available. star α (2000) δ (2000) πH (mas) DH(kpc) DM(kpc) Dg(kpc) l b HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 HD 106191 HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 HD 210709 HD 210910 HD 222349 BD+18 5215 HD 223938 00:11:38 00:28:20 00:55:44 01:21:10 02:01:23 02:10:09 03:37:55 04:18:34 06:44:55 08:54:01 10:02:01 10:22:22 10:40:58 12:13:11 12:21:52 13:26:38 14:17:33 14:09:36 16:21:52 16:44:44 19:59:58 22:12:54 22:13:42 23:40:01 23:46:56 23:53:50 -49:39:20 +10:11:23 -27:53:36 -47:31:48 -04:48:10 -60:45:31 -06:58:25 +02:28:17 +20:51:38 -26:54:56 -33:41:11 -29:33:23 -12:11:44 -15:13:56 -18:24:00 -04:26:46 -83:32:52 -44:22:01 +27:22:27 -25:12:59 -48:58:32 -35:25:51 -03:46:31 -56:44:26 +19:28:22 -50:00:00 7.11(1.08) 27.51(0.86) 0.22(1.42) 13.43(1.16) ... 0.141(20) 0.036(1) 4.5(20.0) 0.074(6) ... 0.161(23) ... 0.979(344) 0.078(7) ... ... ... ... ... 0.219(30) 8.91(0.88) 0.112(10) 0.118(12) ... ... 6.01(1.13) 21.77(1.07) 3.37(1.11) 7.90(1.39) 23.42(0.93) 7.82(1.05) ... 5.02(1.06) 0.23(1.34) 1.73(0.86) 8.75(1.39) 16.64(0.99) ... 14.06(1.18) -0.10(1.37) 6.13(1.16) ... ... ... 0.166(30) 0.046(2) 0.297(98) 0.127(22) 0.043(2) 0.128(17) ... 0.199(42) 4.3(25.3) 0.578(287) 0.114(18) 0.060(4) ... 0.071(6) ... 0.163(31) ... ... ... 0.249(20) 0.168(15) 0.046(2) 0.208(32) 0.160(27) 0.043(2) 0.120(13) ... ... 0.703(172) 0.834(295) 0.121(11) ... ... 0.074(6) 0.197(29) 0.195(32) ... ... ... ... ... ... 0.145(20) ... ... ... ... ... 0.74(10) ... ... ... ... ... 0.168(20) 0.153(20) 4.42(1.21) 0.226(62) 0.218(36) ... 319.03 113.62 251.97 288.97 162.77 286.75 193.68 190.74 193.53 251.43 267.02 268.02 259.85 289.43 293.17 319.33 305.46 317.36 45.97 355.20 350.01 9.08 57.73 321.31 102.68 324.82 -66.21 -52.26 -88.78 -68.78 -62.14 -53.83 -45.71 -32.02 + 7.97 +11.42 +17.12 +23.03 +39.52 +46.63 +43.91 +57.30 -21.10 +16.30 +43.60 +13.21 -31.07 -55.37 -45.72 -57.77 -40.85 -64.61 Table 3. Reddening values for sample stars (see Sect. 3.1). star HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 HD 106191 HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 HD 210709 HD 210910 HD 222349 BD+18 5215 HD 223938 Av 0.0(1) 0.0(1) 0.0(1) 0.0(1) 0.0(1) 0.0(1) 0.030(108) 0.092(123) 0.0(1) 0.038(109) 0.161(140) 0.0(1) 0.045(111) 0.018(105) 0.037(109) 0.0(1) 0.061(115) 0.130(132) 0.0(1) 0.090(123) 0.051(113) 0.0(1) 0.026(106) 0.0(1) 0.047(112) 0.0(1) E(B-V) E(V-I) E(V-R) E(R-I) E(V-K) ... ... ... ... ... ... 0.010 0.029 ... 0.012 0.052 ... 0.014 0.006 0.012 ... 0.020 0.042 ... 0.029 0.016 ... 0.008 ... 0.015 ... ... ... ... ... ... ... 0.012 0.036 ... 0.015 0.063 ... 0.018 0.007 0.015 ... 0.024 0.051 ... 0.035 0.020 ... 0.010 ... 0.019 ... ... ... ... ... ... ... 0.005 0.015 ... 0.006 0.027 ... 0.007 0.003 0.006 ... 0.010 0.022 ... 0.015 0.008 ... 0.004 ... 0.008 ... ... ... ... ... ... ... 0.007 0.021 ... 0.009 0.037 ... 0.010 0.005 0.009 ... 0.014 0.030 ... 0.021 0.012 ... 0.006 ... 0.011 ... ... ... ... ... ... ... 0.027 0.082 ... 0.034 0.143 ... 0.040 0.016 0.033 ... 0.054 0.115 ... 0.080 0.045 ... 0.023 ... 0.042 ... 3.2. Temperatures Literature photometric data were taken from the 2MASS Point Source Catalog (Cutri et al. 2003, Ks magnitude), The General Catalogue Photometric Data by J.C. Mermilliod, B. Hauck and M. Mermilliod available at the web ad- dress http://obswww.unige.ch/gcpd/gcpd.html (B1, B2, G and V1 magnitudes from Geneva system), Hipparcos Catalogue (Perryman et al. 1997, (B-V)H and V, shown in the columns 3 and 8 of Table 4, respectively). For sample stars missing in the Hipparcos Catalogue, (B-V) values were taken from SIMBAD. Photometric data used are shown in Table 4. The colour-temperature calibrations were adopted from Alonso et al. (1996) for stars of log g > 3.6, Alonso et al. (1999) for log g ≤ 3.6 and Lejeune et al. (1998) for all stars. Alonso et al. calibrations use the Johnson system for UBVRI and TCS (Telesc´opio Carlos S´anchez) for JHKLHMN. Lejeune et al. used the Cousins system for UBVRI and MSO for JHKLHMN. The calibrations used for the Geneva colours were adopted from Mel´endez & Ramirez (2003) for dwarfs and Ram´ırez & Mel´endez (2004) for giants. The transfor- mations between photometric systems were obtained from Carpenter (2001), Bessell & Brett (1988), Alonso et al. (1996) and Alonso et al. (1999). The G band affects the spectrum in the λ < 4320Å region. It is possible to see some alteration in the black body distri- bution in the far infrared, probably due to dust in the binary system (Catchpole et al. 1977; Hakkila 1989). The colour (B- V) is affected by the CN and C2 bands, causing in the spectrum the depression of Bond & Neff (1969), making the stars red- 4 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars der and as a consequence the temperatures derived from this colour can be lower. Colours (R-I) and (V-I) are more affected by molecular bands and blanketing effects in K and M stars than (V-K). The calibrations resulted in slightly different tem- peratures. In general, Lejeune et al. (1998) provide higher tem- peratures than Alonso et al. (1996) and Alonso et al. (1999). The resulting temperatures of the Geneva system are generally higher than those from the (B-V) index (see Tables 6 and 7). Given that some indicators tend to increase and others to re- duce the temperature, effective temperatures adopted were the mean values obtained from the (B-V) Hipparcos, (B-V), (V- I), (V-R) and (R-I) from LNA, and (V-K) 2MASS indices and the colours of the Geneva system (B2-V1), (B2-G) and (B1-B2). This is shown in column 2 of Table 9. The uncertainties were estimated to be ±100 K, taking into account the errors from the colour-temperature calibrations. In order to check photometric temperatures, the excitation temperatures Texc were derived by imposing excitation equi- librium. Lines of Fe  and Fe  that give metallicities differing between the average and median of ∆[Fe/H] > 0.01 dex were eliminated. This procedure prevents blended lines to affect the results. Resulting Texc are given in column 3 of Table 9, and illustrated in Figure 1. -2 0 2 4 6 8 10 4 3.9 3.8 3.7 3.6 3.5 3.4 Fig. 2. Isochrones from Bertelli et al. (1994) with sample stars overplotted (black squares). column 4 of Table 9. The gravity log g was varied until the two curves of growth give a same [Fe/H] value, where Fe  lines are more sensitive to gravity variations. According to Nissen et al. (1997), this method presents a problem because Fe  lines are affected by NLTE effects. The uncertainty was estimated as ±0.1 dex. 2) The classical equations of stellar evolution (equation 1) as a function of the distances, according to Nissen et al. (1997) and Allende Prieto et al. (1999): log g g ∗ ⊙! = log M ∗M ⊙! + 4 log Te f f∗ +2 log 1 Te f f⊙! + 0.4V D + 0.1 ◦ + 0.4BC + σlogg = " σM M ln(10)!2 + (0.4σV◦ + 4σT e f f∗ Te f f∗ ln(10)!2 ln(10)!2 + 4σT e f f⊙ π ln(10)!2#0.5 )2 + (0.4σBC)2 + 2σπ Te f f⊙ + + σ2 logg⊙ (1) (2) ◦ is the stellar mass, V is the V corrected magni- where M ∗ tude, BC is the bolometric correction and D is the distance de- rived as explained in Sect. 3.1. For the Sun, Te f f⊙ = 5781 K (Bessell et al. 1998); log g ⊙ = 4.44; Mbol⊙ = 4.75 (Cram 1999) were adopted. adopted from Bertelli et al. isochrones, as shown in Figure 2, and reported in column 9 of Table 10, corresponding to metallicities and temperatures as close as possible to those of the sample (columns 7 (or 6) and 2 of Table 9, respectively). in equation 1 were Stellar masses (1994) Fig. 1. [FeI/H] vs. χex (eV) for the star HD 123585. Least- squares fits for 3 values of temperatures were used to derive Texc. Uncertainties on [FeI/H] and χex were estimated to be ±0.10 and ±0.01, respectively. 3.3. SurfaceGravity Surface gravities (log g) were determined using two methods: 1) Spectroscopic gravities were derived by imposing ion- ization equilibrium of Fe  and Fe  (Figure 3), and given in D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 5 Fig. 3. Fe  and Fe  curves of growth for the star HD 89948. 'W' are the equivalent widths. log(a(sun)) = log(nx/nH), where nx and nH are the numerical densities of an element and hydrogen, respectively; log g f : oscillator strength; Γ = (W/λ)(nH/nx)(g f )−1, as defined in Cayrel & Jugaku (1963) and Spite & Spite (1975). The iron abundance is given by the horizontal distance between the linear part of the curve of growth and the 45◦ line through the origin. Uncertainties were estimated as ± 0.1 M are shown in column 5 of Table 9. ⊙ . These log g values adopted in this work. The lines for which hfs were used were checked using the solar spectrum (Kurucz et al. 1984). According to Nissen et al. (1997), the gravities resulting from the first method are systematically lower than those of the second, but there is no clear explanation. It could reside in NLTE effects in Fe  lines, uncertainties on masses or tempera- tures. 3.4. OscillatorStrengths The Fe  line list and respective oscillator strengths from National Institute of Standards & Technology (NIST) library (Martin et al. 1988, 2002) and Fe  oscillator strengths renor- malized by Mel´endez & Barbuy (2006) were adopted. The list of Fe  and Fe  lines is given in Table 5. Damping constants for neutral lines were computed using the collisional broadening theory of Barklem et al. (1998, 2000, and references therein), as described in Zoccali et al. (2004) and Coelho et al. (2005). The oscillator strengths for the elements other than Fe and re- spective sources are shown in Table 15. For α- and iron peak elements, most values are from NIST. Laboratory values were preferred over theoretical ones. For lines of Cu I, Eu II, La II, Ba II and Pb I, hyperfine structure (hfs) was taken into account employing a code made available by McWilliam, following the calculations de- scribed by Prochaska et al. (2000). The hfs constants were taken from Rutten (1978) for Ba II, Lawler et al. (2001a) for La II, Lawler et al. (2001b) for Eu II and Bi´emont et al. (2000) for Pb I. The final hfs components were determined by using the solar isotopic mix by Lodders (2003) and total log gf values from laboratory measurements, as shown in Table 15. For cop- per, the hfs from Biehl (1976) was used, with isotopic fractions of 0.69 for 63Cu and 0.31 for 65Cu. In this case, small correc- tions were applied such that the total log gf equals the gf-value Table 5. Equivalent widths and atomic constants for Fe  and Fe  lines. e1 - HD 749; e2 - HR 107; e3 - HD 5424; e4 - HD 8270; e5 - HD 12392; e6 - HD 13551; e7 - HD 22589; e8 - HD 27271; e9 - HD 48565; e10 - HD 76225; e11 - HD 87080; e12 - HD 89948; e13 - HD 92545; e14 - HD 106191; e15 - HD 107574; e16 - HD 116869; e17 - HD 123396; e18 - HD 123585; e19 - HD 147609; e20 - HD 150862; e21 - HD 188985; e22 - HD 210709; e23 - HD 210910; e24 - HD 222349; e25 - BD+18 5215; e26 - HD 223938. Full table is only available in electronic form. ion λ χex log gf e1 e2 e3 e4 e5 e6 e7 e8 e9 e10 ... Fe I 4514.19 3.05 -2.050 89 24 76 34 79 36 67 87 24 Fe I 4551.65 3.94 -2.060 48 7 5 ... Fe I 4554.46 2.86 -3.050 ... 13 Fe I 4579.82 3.07 -2.830 47 ... 41 ... Fe I 4587.13 3.57 -1.780 83 26 79 27 80 34 61 86 23 Fe I 4613.21 3.29 -1.670 87 ... 95 46 91 62 73 93 33 . . 41 12 45 ... ... 16 ... 36 ... 31 50 ... ... ... ... 25 ... . . . . . . . . . . . . . . . . . . . . . . . . 32 ... ... ... 9 ... 34 ... 47 ... 3.5. MetallicitiesandMicroturbulentVelocities Equivalent widths were measured with IRAF, and Fe  and Fe  lines with 10 < Wλ < 160 mÅ were considered. Photospheric 1D models were extracted from the NMARCS grid (Plez et al. 1992), originally developed by Bell et al. (1976) and Gustafsson et al. (1975) for gravities log g < 3.3. For less evolved stars, with log g ≥ 3.3 the models by Edvardsson et al. (1993) were adopted. 6 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Table 4. Colours and magnitudes. V magnitudes are from the Hipparcos database, otherwise from SIMBAD (marked with *); Ks from 2MASS Point Source Catalog; B1, B2, V1 and G are Geneva magnitudes; "H": Hipparcos; "S": SIMBAD; "L": LNA. Uncertainties for Hipparcos database are of ± 0.01 whereas for SIMBAD they were estimated in ± 0.05. B2 star (B-V)H (B-V)L (B-V)S V1 Ks G (V-I)L (V-R)L (R-I)L V B1 HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 HD 106191 HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 HD 210709 HD 210910 HD 222349 BD+18 5215 HD 223938 1.130 0.430 1.130 0.490 1.060 0.550 0.710 1.000 0.528 0.510 0.780 0.500 0.470 0.680 0.400 0.970 1.180 0.520 0.470 0.490 0.480 1.060 1.100 0.500 0.370 0.890 1.126(12) 0.447(5) 1.141(2) 0.537(17) ... 0.599(24) ... 1.011(17) 0.562(22) 0.567(22) 0.775(12) 0.545(11) 0.503(3) ... 0.441(8) 1.040(15) 1.190(15) 0.519(12) 0.600(15) ... 0.532(17) 1.103(5) 1.100(30) ... ... 1.126 0.408 1.169 0.539 1.126 0.522 0.752 1.028 ... 0.522 0.761 0.543 0.500 0.600 0.434 1.041 1.224 0.505 0.584 0.515 0.550 1.117 1.086 0.495 ... 1.096 0.506 1.003 0.664 0.842 0.682 0.806 1.051 ... 0.594 0.797 0.612 0.592 0.643 0.524 0.997 1.178 0.555 0.655 0.578 0.604 1.051 1.133 0.571 ... 0.581 0.245 0.521 0.334 0.460 0.351 0.419 0.542 ... 0.300 0.405 0.309 0.299 0.323 0.263 0.520 0.598 0.285 0.332 0.296 0.300 0.559 0.596 0.287 ... 0.515 0.261 0.482 0.330 0.382 0.331 0.387 0.509 ... 0.294 0.392 0.303 0.293 0.320 0.261 0.477 0.580 0.270 0.323 0.282 0.304 0.492 0.537 0.284 ... 0.905(14) 0.873 0.893 0.461 0.432 7.91 6.05 9.48 8.82 8.49* 9.32 8.97* 7.53 7.20 9.23 9.40 7.50 8.55 10.00* 8.55 9.49 8.97 9.28 8.57 9.17* 8.55 9.23 8.49 9.20* 9.74* 8.62 5.385(9) 4.942(13) 7.015(15) 7.524(15) 6.231(31) ... 0.992 1.372 1.013 ... ... 1.401 1.117 1.352 ... ... 1.172 0.283 1.063 ... 1.525 0.466 1.390 ... 1.014 1.363 1.076 1.420 7.265(11) 5.270(11) 5.806(17) 7.979(39) 7.567(19) 6.189(19) 7.282(23) 8.586(21) 7.415(17) 7.143(19) 6.144(17) 8.081(19) ... 8.009(27) 7.231(23) 6.711(19) 5.857(19) 7.940(25) 8.535(11) 6.504(13) ... ... 1.007 1.017 1.120 1.036 1.016 1.053 0.983 ... ... 1.015 1.010 1.020 1.034 ... 1.307 1.001 0.991 ... ... ... 1.355 1.359 1.275 1.338 1.362 1.330 1.384 ... ... 1.362 1.359 1.355 1.345 ... 1.144 1.365 1.368 ... ... ... 1.028 1.058 0.769 1.021 1.080 0.966 1.153 ... ... 1.061 1.078 1.068 1.019 ... 0.344 1.082 1.110 ... 1.352 1.390 1.050 1.347 1.412 1.284 1.505 1.405 1.421 1.399 1.347 0.513 1.421 1.452 Microturbulent velocities vt were determined by canceling the trend of Fe  abundance vs. equivalent width. The uncertainties on the metallicity are due to uncertainties on the input parameters: temperature, log g and microturbu- lent velocity. The variation of these parameters affects the iron abundance ǫ(Fe), as follows σǫ(Fe) = q(∆ǫ(Fe)T )2 + (∆ǫ(Fe)lg)2 + (∆ǫ(Fe)v)2 (3) where ∆ǫ(Fe)T , ∆ǫ(Fe)lg and ∆ǫ(Fe)v are the differences on the iron abundances due to variations on temperature, log g and microturbulent velocity, respectively. The contributions of the equivalent widths (from 0.5 to 0.9 mÅ) to the uncertainty in metallicity are negligible (Cayrel 1989), where uncertainties on continuum placement are not taken into account. For HD 5424, a variation of ∆T = 100 K in the tempera- ture results in ∆ǫ(Fe)T = 1.940 × 106, ∆ log g = 0.3 dex in log g results in ∆ǫ(Fe)lg = 2.855 × 106 and ∆vt = 0.1 km/s in vt results in ∆ǫ(Fe)v = 0.803 × 106 in the ǫ(Fe). For HD 150862, ∆ǫ(Fe)T = 5.65 × 105 for ∆T = 100 K, ∆ǫ(Fe)lg = 1.69 × 106 for ∆ log g = 0.1 dex and ∆ǫ(Fe)v = 0.113 × 106 for ∆vt = 0.1 km/s. The quantity of interest is the logarithm of ǫ(Fe), such that σlog ǫ(Fe) = (σǫ(Fe))/(ǫ(Fe) ln 10). Consequently, σ[Fe/H] = , resulting σ[Fe/H] = 0.18 for HD 5424, and this result should be typical for all sample stars with log g < log ǫ(Fe) + σ2 qσ2 log ǫ(Fe)⊙ 3.3. For HD 150862, σ[Fe/H] = 0.04, this uncertainty being adopted for stars with log g ≥ 3.3. 3.6. AdoptedAtmosphericParameters The atmospheric parameters were derived in an iterative way, adopting initial values of log g and [Fe/H] according to North et al. (1994) and G´omez et al. (1997). The lines of Fe  and Fe  were used separately in order to test the ionization equilibrium (see Sect. 3.3). In this first iteration, one searchs for values of log g, [Fe/H] and vt corresponding to the ionization equilibrium in order to verify the trend of the results. Using the average of the temperatures and masses from isochrones, the surface gravity was determined from equation 1, in order to be compared with that from the ionization equilibrium. In the first iteration, values different from the ones adopted initially for [Fe/H] and log g from equation 1 were obtained. In this case, these new values were used as input in the colour- temperature calibrations and the procedure was restarted. The procedure was repeated until the input parameters matched the output ones, providing a consistent set of atmospheric param- eters. After several iterations, it was possible to define a set of atmospheric parameters for each star, shown in Table 9. These results indicate that our sample contains giants, subgiants and dwarfs. Table 11 shows atmospheric parameters for barium stars found in the literature. D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 7 Table 7. Temperatures based on calibrations by Alonso et al. (1996) or Alonso et al. (1999) and for the Geneva system those by Mel´endez & Ramirez (2003) for subgiants or dwarfs and Ram´ırez & Mel´endez (2004) for giants. star HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 HD 106191 HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 HD 210709 HD 210910 HD 222349 BD+18 5215 HD 223938 T(B−V)S 4570 6440 4430 6130 4670 5890 5420 4950 5910 6150 5280 6150 6390 5550 6520 4760 4230 6160 6400 6350 6300 4670 4660 6020 6670 4930 T(B−V)H 4570 6370 4410 5940 ... 5730 ... 4930 5790 5930 5290 5970 6250 ... 6330 4620 4220 6170 5880 ... 6080 4600 4660 ... ... 4900 T(B−V)L 4570 6540 4360 5940 4560 6000 5290 4890 ... 6110 5330 5980 6260 5820 6360 4620 4160 6220 5940 6240 6010 4570 4680 6040 ... 4970 T(V−I)L 4530 6460 4720 5760 5110 5700 5320 4700 ... 6110 5510 5970 6130 5870 6450 4740 4430 6470 5800 6280 6090 4620 4480 6150 ... 4980 T(V−R)L 4570 6600 4740 5860 5010 5740 5370 4780 ... 6180 5560 6070 6240 6000 6440 4750 4480 6400 5950 6310 6210 4630 4540 6190 ... 4990 T(R−I)L 4570 6180 4690 5650 5190 5640 5300 4690 ... 5970 5480 5820 5950 5720 6330 4710 4360 6430 5620 6150 5910 4660 4510 6040 ... 4920 T(V−K) 4590 6410 4620 6080 4840 ... 5470 4940 5920 6210 5450 6060 6190 5920 6430 4740 4360 6460 ... 6460 6120 4590 4520 6140 6310 4990 ... TB2−V1 6410 4400 6050 ... 6060 ... ... 5820 6070 5250 5950 6220 5780 6370 ... ... 6140 6210 6260 5960 ... 4520 6040 6270 ... TB2−G ... 6430 4470 6050 ... 6120 ... ... 5860 6110 5370 5990 6240 5850 6380 ... ... 6220 6280 6280 6020 ... 4570 6060 6280 ... ... TB1−B2 6360 4460 6050 ... 6130 ... ... 5820 6080 5370 6000 6200 5830 6370 ... ... 6240 6210 6270 6020 ... 4430 6070 6280 ... The use of models by Gustafsson et al. (1975) led to ∆[FeI/H] and ∆[FeII/H] ≤ 0.03 dex relative to models by Edvardsson et al. (1993) and Plez et al. (1992) for most stars. For two stars, ∆[Fe/H] ≈ 0.15 dex and for another 4 ones, 0.04 dex < ∆[Fe/H] < 0.09 dex. In Table 10 are given the bolometric corrections BC(V) de- termined according to Alonso et al. (1995) for dwarfs and sub- giants and to Alonso et al. (1999) for giants, with good agree- ment with values determined from a linear interpolation on Lejeune et al. (1998) grids. The uncertainties on BC(V) were estimated by computing how the uncertainties on temperatures modify its value; from equations 4 to 12, one obtains the ab- solute magnitude Mv, the bolometric magnitude Mbol, the lu- minosity (L ) and the mass (M ), providing as input the distance D(pc), Av, V and log g: ), the radius (R ∗/R ⊙ ∗/M ∗/L ⊙ ⊙ Mv = V − 5 log D + 5 − Av D ln(10)!2 σMv = "(σV)2 + 5σD Mbol∗ = Mv + BC(V) σMbol∗ = σ2 ∗ = 10−0.4(Mbol∗−Mbol⊙ L BC!0.5 Mv + σ2 )L ⊙ + σ2 Av#0.5 (4) (5) (6) (7) (8) R Mbol⊙!0.5 Mbol∗ + σ2 ∗ L ⊙ σL = L0.4 ln(10) σ2 T 4 !0.5 ∗ = L e f⊙ T 4 e f∗ 2L!2 Te f f⊙ !2 + 2σT e f f⊙ σR = R" σL Recalling that g = 10log(g) and σg = g ln(10)σlog(g) Te f f∗ !2#0.5 + 2σT e f f∗ R ⊙ M ∗ = g g ∗ ⊙ R2 ∗ R2 ⊙ M ⊙ g∗ !2 σM = M" σg∗ g⊙ !2 + σg⊙ + 2σR R !2#0.5 . (9) (10) (11) (12) (13) For masses computed with equation 12 and log g from ion- ization equilibrium, the results are those shown in column 8 of Table 10. These masses are lower than those derived from the isochrones (column 9), resulting in higher values for log g computed from equation 1. are compatible with Mennessier et al. (1997) masses for barium stars, accord- ing to their luminosity classes: 1 - 1.6 M for dwarfs, and 1 - 3 M for giants and subgiants. For stars with no Hipparcos parallaxes, distances also had to be derived iteratively. In the cases where extinction was null, derivations essentially Our ⊙ ⊙ 8 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Table 6. Temperatures based on Lejeune et al. (1998) calibra- tions. Table 8. Masses and distances adopted for stars with no Hipparcos parallax πH values. ... ... ... ... 5720 ... ... 4710 6400 4520 5960 4990 5760 5930 5370 5970 6290 4770 6630 4530 5990 4610 6030 5390 5010 4570 6450 4690 5870 5230 5800 5420 4700 4590 6540 4660 5880 5100 5760 5380 4800 T(B−V)S T(B−V)H T(B−V)L T(V−I)L T(V−R)L T(R−I)L T(V−K) star 4540 4620 4700 HD 749 6370 6460 6480 HR 107 4700 4650 4550 HD 5424 6160 5860 6200 HD 8270 5430 4870 4690 HD 12392 5850 5890 HD 13551 5480 5600 5460 HD 22589 4630 4950 5010 HD 27271 5930 6080 HD 48565 6180 6310 6170 HD 76225 5680 5400 5360 HD 87080 6050 6180 6160 HD 89948 6440 6180 6270 HD 92545 5940 6050 5570 HD 106191 6450 6460 6550 HD 107574 4660 4760 4910 HD 116869 4430 4380 4410 HD 123396 6210 6560 6500 HD 123585 5870 6450 HD 147609 6370 6500 6410 HD 150862 6140 6250 6360 HD 188985 4610 4620 4780 HD 210709 4740 4500 4550 HD 210910 6240 6270 HD 222349 6040 BD+18 5215 6800 6400 4890 5000 HD 223938 5000 6170 5600 6060 6220 6000 6390 4690 4450 6390 5970 6300 6210 4620 4540 6190 6180 5630 6060 6200 5970 6420 4680 4440 6460 5920 6330 6170 4620 4520 6220 6150 5440 6010 6350 5860 6410 4820 4310 6330 6000 6340 6080 4720 4830 6100 4970 5080 4940 4980 ... 6340 4780 4350 6220 5890 ... 6120 4690 4740 ... ... ... ... ... ... ... ... the determination of the colours was straight forward, the tem- peratures and the other parameters, log g, [Fe/H] and BC. The distance can be determined by setting the mass. In the case of extinction variation on which an average of Av was adopted, the distance obtained can be different from that used by com- puting Av. In this case, the new distance was used to restart the procedure, computing new Av, colours and temperatures. The procedure was repeated until the output distance matched the input one. In both cases, the calculation was made for several masses inside the range given by Mennessier et al. (1997) and the absolute magnitudes were used to derive the masses from isochrones in order to be compared with the input masses. The log g was obtained with equation 1. The mass chosen was that for which log g from the equation 1 was inside the range of 0.3 dex of the spectroscopic one. Once the mass was set, the distances were recalculated, as shown in Table 8. According to Th´evenin & Idiart (1999), the NLTE affects Fe  lines more than Fe  ones, and it directly affects the log g derived from ionization equilibrium. For this reason, the grav- ities adopted were those resulting from expression 1, whereas for the metallicities, Fe  lines were used, as shown in column 7 of Table 9. It is worth noting that using the gf-values for Fe  lines by Mel´endez & Barbuy (2006), the abundances of Fe de- rived from the Fe  and Fe  lines are closer to each other and the log g from ionization equilibrium increases by 0.2 dex rel- ative to the case of the gf-values from NIST. star Dmin-Dmax Mmin-Mmax adopted HD 12392 HD 22589 HD 106191 HD 150862 HD 222349 BD+18 5215 150-240pc 240-288pc 140-180pc 70-90pc 154-182pc 140-190pc 1.6-2,4M ⊙ 1,6-1,8M ⊙ 1,0-1,1M ⊙ 1,0-1,2M ⊙ 1,1-1,2M ⊙ 1,1-1,2M ⊙ 219(30)pc 2,0(1)M ⊙ 249(20)pc 1,6(1)M ⊙ 145(20)pc 1,0(1)M ⊙ 74(10)pc 1,1(1)M ⊙ 168(20)pc 1,2(1)M ⊙ 153(20)pc 1,1(1)M ⊙ Table 9. Stellar parameter results. Te f f : photometric tem- perature; Texc: excitation temperature; logg(C): logg related to curve of growth; logg(D): logg related to masses from isochrones. Numbers in parenthesis are errors in last decimals. star Te f f Texc HD 749 4610 4580 HR 107 6440 6650 HD 5424 4570 4700 HD 8270 5940 6070 HD 12392 5000 5000 HD 13551 5870 6050 HD 22589 5400 5630 HD 27271 4830 4830 HD 48565 5860 6050 HD 76225 6110 6330 HD 87080 5460 5550 HD 89948 6010 6010 HD 92545 6210 6270 HD 106191 5890 5890 HD 107574 6400 6400 HD 116869 4720 4850 HD 123396 4360 4480 HD 123585 6350 6450 HD 147609 5960 5960 HD 150862 6310 6310 HD 188985 6090 6190 HD 210709 4630 4680 HD 210910 4570 4770 HD 222349 6130 6190 BD+18 5215 6300 6300 HD 223938 4970 5150 logg(C) logg(D) [Fe /H] [Fe /H] vt +0.17 0.9 1.6 -0.36 1.1 -0.55 0.9 -0.42 -0.12 1.2 1.1 -0.24 -0.27 1.1 +0.17 1.3 1.0 -0.62 -0.31 1.4 1.0 -0.44 1.2 -0.30 1.3 -0.12 1.1 -0.29 -0.55 1.6 1.3 -0.32 1.2 -0.99 -0.48 1.7 +0.08 1.5 1.4 -0.10 1.1 -0.30 -0.04 1.1 +0.04 2.0 -0.63 1.1 1.5 -0.53 -0.13 1.0 2.8(1) 4.08(7) 2.0(3) 4.2(1) 3.2(1) 4.0(1) 3.3(1) 2.9(1) 4.01(8) 3.8(1) 3.7(2) 4.30(8) 4.0(1) 4.2(1) 3.6(2) 2.2(2) 1.4(3) 4.2(1) 4.42(9) 4.6(1) 4.3(1) 2.4(2) 2.7(2) 3.9(1) 4.2(1) 3.1(1) -0.06 -0.34 -0.51 -0.44 -0.06 -0.44 -0.12 -0.09 -0.71 -0.34 -0.49 -0.28 -0.15 -0.22 -0.56 -0.35 -1.19 -0.44 -0.45 -0.11 -0.25 -0.07 -0.37 -0.58 -0.44 -0.35 2.3 4.0 1.8 4.2 3.2 3.7 3.3 2.3 3.8 3.7 3.7 4.2 4.0 4.2 3.6 2.1 1.2 4.2 3.3 4.6 4.3 2.3 2.0 3.9 4.2 2.7 from the colours indices in the present work for HD 27271 and HD 147609 are lower than those found in the literature. Furthermore, the metallicities used were those derived from Fe  lines instead of ionization equilibrium. Also the masses derived using log g from ionization equilibrium are very small (see column 8 of Table 10), therefore we preferred to use log g derived by using masses from isochrones, and Fe  lines for the determination of metallicities. 4. Abundances Table 9 compared with Table 11, shows that differences are stronger for HD 27271 and HD 147609. Temperatures derived The LTE abundance analysis and the spectrum synthesis cal- culations were performed using the codes by Spite (1967, and D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 9 Table 10. Bolometric corrections, absolute magnitudes, limi- nosities, radii and masses for the sample stars. BCa(V): bolo- metric corrections using Alonso et al. (1996) for dwarfs and subgiants and Alonso et al. (1999) for giants; BCl(V): bolo- metric corrections using Lejeune et al. (1998); M : stel- lar masses from curve of growth; Mi/M : stellar masses from ⊙ isochrones. Numbers in parenthesis are errors in last decimals. ∗/M ⊙ star BCa(V) BCl(V) Mv Mbol M ∗/M R ∗/R ⊙ ⊙ HD 749 0.4(1) 7(1) -0.43(6) -0.47(7) 1.9(3) 1.4(3) HR 107 1.0(3) -0.08(1) -0.04(1) 3.3(1) 3.2(1) 1.6(1) HD 5424 1.1(8) -0.44(6) -0.48(7) -0.5(8) -0.9(8) 184(130) 22(8) HD 8270 0.8(3) 1.2(1) -0.13(1) -0.12(1) 4.4(2) 4.2(2) HD 12392 2.0(8) -0.26(3) -0.28(2) 1.8(3) 1.5(3) 6.0(9) HD 13551 0.4(1) 1.5(2) -0.14(1) -0.12(1) 4.0(2) 3.8(2) HD 22589 1.4(5) 4.5(6) -0.20(2) -0.19(2) 2.0(3) 1.8(3) HD 27271 0.5(2) 8(1) -0.32(4) -0.35(5) 1.3(2) 1.0(2) HD 48565 0.5(1) 1.5(1) -0.15(1) -0.14(1) 3.9(1) 3.7(1) HD 76225 1.1(5) 2.5(4) -0.08(1) -0.09(1) 2.6(3) 2.5(3) HD 87080 1.1(3) -0.20(2) -0.19(2) 3.2(4) 3.0(4) 2.5(5) HD 89948 1.18(9) 0.8(2) -0.11(1) -0.10(1) 4.3(1) 4.2(1) HD 92545 1.3(5) 1.9(2) -0.08(1) -0.06(1) 3.1(3) 3.0(3) HD 106191 1.0(4) 1.3(2) -0.11(1) -0.11(1) 4.2(3) 4.0(3) HD 107574 1.3(6) 3.0(6) -0.08(1) -0.06(1) 2.0(5) 1.9(5) HD 116869 0.9(5) -0.37(4) -0.38(5) 0.2(5) -0.1(5) 14(4) HD 123396 -0.57(7) -0.61(7) -0.7(8) -1.3(8) 255(180) 28(10) 0.4(3) HD 123585 1.4(2) -0.10(4) -0.06(1) 3.7(2) 3.6(2) 1.1(4) HD 147609 1.02(9) 0.07(2) -0.10(1) -0.09(1) 4.7(2) 4.6(2) HD 150862 1.1(4) 0.9(1) -0.07(1) -0.06(1) 4.7(3) 4.7(3) HD 188985 1.1(2) 1.2(1) -0.10(1) -0.09(1) 4.1(2) 4.0(2) HD 210709 0.8(4) 11(2) -0.53(6) -0.45(7) 1.1(4) 0.5(4) 0.2(1) 7(1) HD 210910 -0.45(6) -0.50(7) 2.0(4) 1.6(4) HD 222349 -0.12(1) -0.11(1) 3.1(2) 2.9(2) 2.0(2) 1.2(4) 1.1(4) 1.4(2) BD+18 5215 -0.10(1) -0.08(1) 3.8(3) 3.7(3) HD 223938 -0.31(4) -0.28(2) 1.9(4) 1.6(4) 6(1) 0.6(2) 1.6(3) 20(6) 2.3(5) 16(4) 32(7) 2.5(3) 8(2) 5(2) 1.6(2) 5(1) 1.9(6) 13(6) 88(40) L ∗ /L ⊙ 21(6) 4.2(5) 2.8(6) 1.2(2) 1.1(3) 1.9(4) 48(20) 19(6) 5(1) 3(1) 18(6) Mi/M ⊙ 1.2 1.2 1.9 0.9 2.0 0.9 1.6 1.9 0.9 1.4 1.2 1.0 1.3 1.0 1.4 1.2 0.8 1.1 1.0 1.1 1.1 1.1 1.0 1.2 1.1 1.4 subsequent improvements in the last thirty years), described in Cayrel et al. (1991) and Barbuy et al. (2003). Table 15 shows the resulting abundances (log ǫ(X) and [X/Fe]) for all atomic lines, whereas Tables 16 and 17 show the mean abundance of each element, obtained for the 26 sam- ple barium stars. Figures 11 to 14 show [X/Fe] vs. [Fe/H]. For most elements, the solar abundances used were extracted from Grevesse & Sauval (1998), otherwise references are indicated in Table 17. For the stars HD 749, HD 13551, HD 27271, HD 123396, HD 147609, HD 210910 and HD 223938, the difference between the metallicities derived from Fe  and Fe  lines, ∆[Fe/H] = [Fe /H] - [Fe /H] ≥ 0.2 dex (see Table 9). There are several possible explanations for it, such as the NLTE effect in lines of Fe , imprecision in stellar parameters (Te f f , log g, vt), blends in Fe  and Fe  lines. Simmerer et al. (2004) also ob- served this effect in their sample of 159 giants and dwarfs. They established relations between ∆[Fe/H] and Te f f or log g and, despite having a dispersion, it is possible to note that the differ- ences seem to be larger at lower temperatures. Kraft & Ivans (2003) attributed this effect to an inadequacy of model atmo- spheres. Yong et al. (2003) observed a similar behaviour in the relation between ∆[Fe/H] and Te f f for giants of the metal-poor globular cluster NGC 6752, also becoming more pronounced for the cooler stars. For the 7 stars mentioned before, the resulting [X/Fe] were very low, as shown by the starred symbols of Figures 11 to 14, and for this reason, the abundances were also determined by using metallicities from Fe  lines, as shown in Figures 11 to 14 and Tables 15, 16 and 17, and they were used in Figures 4, Table 11. Atmospheric parameters for barium stars col- lected in the literature. References: E93: Edvardsson et al. (1993); G96: Gratton et al. (1996); T99: Th´evenin & Idiart (1999); P05: Pereira (2005); P03: Pereira & Junqueira (2003); B92: Barbuy et al. (1992); N94: North et al. (1994); S86: Smith & Lambert (1986); L91: Luck & Bond (1991); S93: Smith et al. (1993). Numbers in parenthesis are errors in last decimals. star Te f f (K) log g [M/H] vt(km/s) ref HR 107 HR 107 HR 107 HD 8270 HD 13551 HD 22589 HD 27271 HD 48565 HD 48565 HD 76225 HD 87080 HD 89948 HD 89948 HD 89948 HD 92545 HD 106191 HD 107574 HD 123585 HD 123585 HD 147609 HD 147609 HD 150862 HD 150862 HD 188985 HD 222349 BD+18 5215 6488 6431 6462 6100 6400 5600 5350 5910 5929 6010 5600 5929 6000 5950 6240 5840 6340 6047 6000 6270 6300 6135 6200 5960 6000 6290 4.08 4.06 4.17 4.2 4.4 3.8 2.30 3.50 3.72 3.50 4.00 4.10 4.00 4.10 3.90 4.05 3.63 3.50 3.50 3.50 3.61 4.05 4.00 3.78 3.76 4.49 -0.37 -0.37 -0.21 -0.53 -0.28 -0.16 -0.50 -0.90 -0.54 -0.50 -0.51 -0.31 -0.13 -0.27 -0.33 -0.40 -0.80 -0.50 -0.50 -0.50 -0.36 -0.30 -0.22 -0.30 -0.90 -0.50 - 1.98 1.00 1.4 1.6 1.4 3.0 1.2(2) 1.00 1.7(2) 1.2 1.00 1.80 0.80 1.7(1) 1.0(1) 1.9(3) 1.8(3) 2.00 1.2(2) 1.20 1.2(1) 2.20 1.4(1) 1.8 1.2(2) E93 G96 T99 P05 P05 P05 B92 N94 T99 N94 P03 S86 L91 S93 N94 N94 N94 N94 L91 N94 T99 N94 L91 N94 N94 N94 5, 6, 9, 10, 15, 16, 17, 18. For the remaining stars, abundances in these figures were obtained using metallicities derived from Fe  lines. The abundance results of the sample stars are essen- tially homogeneous, even considering the different luminosity classes, as shown in Figures 11 to 14. Regarding Al, Na, α- and iron peak elements, the behaviour of [X/Fe] vs. [Fe/H] is in agreement with disk stars. For heavy elements, there is a variation that could be explained by the amount of enriched material that each star received from the more evolved com- panion. The sample is too small to reveal differences between the 4 halo stars and the disk ones. The overabundance found for the s-elements in the sample stars is expected for barium stars and this peculiarity is independent of the luminosity class. Literature abundance data for the present sample barium stars are shown in Table 20. 10 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Fig. 4. Lithium abundance as a function of temperatures. Symbols: squares: dwarf stars (log g ≥ 3.7); triangles: subgiants (2.4 < log g < 3.7); circles: giants (log g ≤ 2.4). The arrows indicate an upper limit. Fig. 5. [C,N/Fe] and [C/O] as a function of [Fe/H]. Symbols are the same as in Figure 4. The arrows indicate an upper limit. In the dashed region are the disk and halo stars of Goswami & Prantzos (2000). Fig. 6. [PCNO/Fe] as a function of Te f f or log g. Symbols are the same as in Figure 4. 4.1. Lithium Lithium is among the yields of primordial nucleosynthesis. , through It can also be produced in stars with M ≤ 8 M spallation by cosmic rays (Woosley & Weaver 1995) and the ν-process suggested for the first time by Domogatskii et al. (1977) and later by Woosley et al. (1990) and Timmes et al. ⊙ (1995). Red Giant Branch or Asymptotic Giant Branch stars possibly produce Li (Sackmann & Boothroyd 1992, 1999) and some of them become Lithium Rich Giants (Brown et al. 1989; Castilho et al. 1998). However, Li is mainly destroyed during the life of a star, and for this reason, it can be used as an ap- praiser of its age. Cooler stars have a deeper convective zone, D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 11 Table 12. Derived abundances of Li, C and N. The stars codes in the header are: e1 - HD 749; e2 - HR 107*; e3 - HD 5424; e4 - HD8270*; e5 - 12392; e6 - 13551*; e7 - HD 22589; e8 - HD 27271; e9 - HD 48565*; e10 - HD 76225*; e11 - HD 87080*; e12 - HD 89948*; e13 - HD 92545*; e14 - HD 106191*; e15 - HD 107574*; e16 - HD 116869; e17 - HD 123396; e18 - HD 123585*; e19 - HD 147609*; e20 - HD 150862*; e21 - HD 188985; e22 - HD 210709; e23 - HD 210910*; e24 - HD 222349*; e25 - BD+18 5215*; e26 - HD 223938. The G band covers the region λλ4295 - 4315 Å. The stars with '*' show an upper limit for N. el mol λ e1 e2 e3 e4 e5 e6 e7 e8 e9 e10 e11 e12 e13 e14 e15 e16 e17 e18 e19 e20 e21 e22 e23 e24 e25 e26 6707.776 -0.33 <1.44 0.10 <1.08 0.58 <1.36 <0.63 0.37 <0.68 <0.94 0.36 1.10 <0.98 <1.11 6707.927 -0.33 <1.44 0.50 <1.38 0.58 <1.12 <0.63 0.37 <0.68 <0.63 0.36 1.10 <0.98 <1.11 ... ... Li Li C C2 5135.600 8.62 C C2 5165.254 8.67 C C2 5635.500 8.67 C CH G band 8.67 N CN 6477.200 7.89 ... N CN 6478.400 N CN 6478.700 ... N CN 6479.000 8.09 N CN 6703.968 7.79 N CN 6706.733 7.79 N CN 6708.993 7.79 N CN 8030.410 7.74 N CN 8030.720 7.79 N CN 8034.970 7.79 N CN 8040.100 7.69 N CN 8040.220 7.69 ... ... ... 8.39 8.35 8.30 8.79 8.39 8.35 8.40 8.84 8.39 8.40 8.50 8.79 8.39 8.35 8.40 8.79 8.66 7.65 8.80 8.50 8.30 8.66 7.65 8.40 8.66 7.95 8.66 8.15 ... 8.66 7.71 8.60 8.45 8.66 7.71 8.60 8.40 8.66 7.71 8.60 8.40 7.71 8.40 8.30 ... 7.81 8.40 8.40 ... 7.75 7.90 8.30 ... 7.55 8.25 8.25 ... ... 7.65 8.25 8.30 8.18 8.28 8.48 8.28 8.75 9.15 8.75 8.75 8.55 8.55 8.75 8.25 ... 7.55 7.95 7.95 ... 8.55 8.59 8.27 8.55 8.69 8.27 8.55 8.69 8.27 8.55 8.69 8.27 8.13 8.35 7.98 7.63 8.35 7.88 7.63 8.35 7.88 ... 7.88 7.93 8.35 7.88 7.93 8.35 7.88 ... 8.35 7.88 7.83 8.15 7.78 7.83 8.25 7.78 7.78 8.35 7.78 7.73 8.20 8.28 7.73 8.25 8.28 8.58 8.58 8.58 8.47 8.74 8.74 8.74 8.74 8.64 8.64 8.64 8.44 8.14 7.94 8.44 8.44 ... ... ... ... ... ... ... ... 8.38 8.47 8.63 8.40 8.47 8.68 8.42 8.78 8.42 8.47 8.78 ... 7.71 ... 7.71 ... 7.71 ... 7.71 8.04 7.71 7.71 8.04 7.71 8.04 7.71 8.10 8.04 7.81 8.10 8.04 7.71 7.85 8.04 7.71 8.05 8.04 7.71 8.05 8.04 8.60 8.37 8.60 8.37 8.65 8.37 8.65 8.32 ... ... ... ... ... ... ... ... 9.31 ... ... 9.31 ... 9.31 8.77 9.31 8.47 9.21 8.07 ... 8.32 9.31 8.32 8.91 8.31 8.31 8.31 8.31 7.93 7.83 7.88 8.08 7.83 7.83 7.83 7.73 7.93 7.93 7.68 7.73 ... <0.18 -0.14 <1.12 <1.18 <1.00 <1.00 <0.16 -0.06 1.47 1.67 0.17 ... <0.18 -0.14 <1.12 <1.18 <1.00 <1.00 <0.16 -0.06 1.77 1.67 0.17 8.42 8.28 8.53 8.53 8.52 8.48 8.53 8.53 8.52 8.68 8.63 8.53 8.32 8.28 8.63 8.53 ... 7.94 8.07 ... ... 8.07 ... ... 8.07 ... ... 8.07 ... 7.94 8.07 ... 7.94 8.07 ... 8.07 7.94 8.79 7.84 7.97 7.97 8.79 7.94 7.97 8.09 8.79 7.84 8.79 7.84 7.97 7.97 8.79 7.84 8.86 8.86 8.86 8.86 8.38 8.38 8.38 8.38 8.38 8.38 8.38 8.38 8.38 8.08 ... ... 8.46 8.46 8.46 8.46 8.13 8.13 8.13 8.13 8.13 8.13 8.13 8.13 8.13 8.13 7.93 8.03 8.61 8.61 8.61 8.61 8.68 8.68 8.68 8.68 8.48 8.48 8.48 8.58 8.18 7.88 8.08 8.08 7.83 7.83 7.81 7.83 ... ... ... ... 6.76 6.76 6.76 6.51 6.75 6.66 6.51 6.61 8.60 8.60 8.60 8.50 8.36 8.36 8.36 8.36 8.36 8.36 8.36 8.36 8.36 ... 8.36 8.36 8.83 8.83 8.83 8.83 8.80 8.80 8.80 ... ... ... ... 8.80 8.80 ... 8.80 8.80 ... ... ... ... ... ... ... ... ... ... ... 1 0.8 0.6 0.4 1 0.8 0.6 0.4 0.2 5162 1 0.8 0.6 0.4 0.2 5135.4 5135.6 5135.8 5136 5163 5164 5165 5166 5634 5634.5 5635 5635.5 5636 1 0.8 0.6 0.4 1 0.8 0.6 0.4 6477 6477.5 6478 6478.5 6479 6479.5 6704 6706 6708 6710 Fig. 7. Example of fits of the C2 bands for the star HD 12392. Symbols: solid line: observed spectrum; dotted lines: synthetic spectra with several C abundances. Upper panel: log ǫ(C) = 8.84, 8.79 e 8.74; Middle panel: log ǫ(C) = 8.89, 8.84 e 8.79; Lower panel: log ǫ(C) = 8.84, 8.79 e 8.74. Fig. 8. Example of fits of CN bands for the star HD 12392. Symbols: solid line: observed spectrum; dotted lines: synthetic spectra with several N abundances. Upper panel: log ǫ(N) = 8.60, 8.50, 8.40, 8.30; Lower panel: log ǫ(N) = 8.50, 8.45, 8.40, 8.35. consequently, Li is brought to inner regions where it is com- pletely destroyed. For low metallicity dwarfs there is an upper limit of the temperature where the convective zone acts such that surface Li can be preserved, forming the Spite's plateau (Spite & Spite 1982). Castilho et al. (2000) showed that the globular cluster NGC 6397 presents a Li dilution curve where for Te f f ≈ 6000 K, log ǫ(Li) ≈ 2.2 and Te f f ≈ 4200 K, log ǫ(Li) ≈ -0.8, indicating an additional Li destruction. Other causes for Li depletion have been suggested such as rotationally induced mixing and mass loss. Most AGB stars rich in O and C are poor in Li (Boesgaard 1970; Denn et al. 1991; Kipper & Wallerstein 1990). Therefore, a barium star considered as a result of the mass transfer from the companion AGB should be also Li-poor. 12 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Table 13. Average abundances from molecular lines and Li synthesis. '<' indicates an upper limit. HD 13551 and HD 22589 from C I lines at λ5380 Å, λ7113 Å, λ7115 Å, λ7117 Å, λ7120 Å, and found, respectively, [C/Fe] = 0.71, 0.46, 0.64 (see Table 20), which values are higher than the present analysis for the same stars (0.31, 0.24, 0.30, Table 13). For the star HD 87080, Pereira & Junqueira (2003) found [C I/Fe] = 0.61, also higher than in this work (0.33), but it is well known that the triplet at λ7115 - λ7120 Å used by them is sub- ject to strong NLTE effects (Przybilla et al. 2001). Comparing the results of [C/Fe] by North et al. (1994) with the stars in common with the present sample shown in Tables 20 and 13, a good agreement is seen, though their results for the stars HD 48565, HD 76225 and HD 107574 are higher. Their results are in the range -0.15 ≤ [C/Fe] ≤ 0.89, confirming the good agree- ment. Figure 5 shows a decreasing trend of [C/Fe] toward higher metallicities for the sample barium stars. Reddy et al. (2003) found similar behaviour for their sample of F and G disk dwarfs, whereas Goswami & Prantzos (2000) found a constant behaviour centered in [C/Fe] ∼ 0 for their sample of halo and disk stars, in the ranges -0.4 < [C/Fe] < +0.4 and -1 < [Fe/H] < 0. It is also possible to note that even the metal-poor barium stars are richer in C than normal stars of similar metallicities. The star HD 123396 does not follow the decreasing trend of [C/Fe] of Figure 5, and, if it is a halo giant, then it is C-rich rel- ative to most stars of the halo and instead compatible with C- rich halo stars (Rossi et al. 2005). Figure 5 also shows that the less evolved stars tend to have higher C abundances than more evolved ones, since they did not reach yet the first dredge-up (see Sect. 4.3). [C/O] vs. [Fe/H] (Figure 5) shows a large dispersion. The range shown is -0.19 ≤ [C/O] ≤ 0.46, and for 6 stars [C/O] < 0 (see Table 13). Barbuy et al. (1992) found the range 0.19 ≤ [C/O] ≤ 0.47. North et al. (1994) found higher values for [C/O], inside the range 0.4 ≤ [C/O] ≤ 1.5. star log ǫ(C) [C/Fe] log ǫ(N) [N/Fe] [C/O] [PCNO/Fe] log ǫ(Li) HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 HD 106191 HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 HD 210709 HD 210910 HD 222349 BD+18 5215 HD 223938 8.66 8.39 8.36 8.41 8.80 8.32 8.55 8.67 8.27 8.56 8.41 8.47 8.72 8.63 8.36 8.31 7.83 8.83 8.58 8.86 8.61 8.46 8.45 8.46 8.58 8.53 0.20 0.23 0.39 0.31 0.40 0.24 0.30 0.24 0.37 0.35 0.33 0.25 0.32 0.40 0.39 0.11 0.50 0.79 0.51 0.44 0.39 -0.02 0.30 0.57 0.59 0.36 7.82 -0.04 -0.01±0.20 <8.66 <1.10 0.30±0.14 7.78 0.41 0.21±0.20 <8.49 <0.99 0.23±0.14 0.57 0.22±0.20 8.37 <8.64 <1.16 -0.19±0.14 7.84 0.19 0.27±0.14 8.31 0.48 0.05±0.20 <7.97 <0.67 0.34±0.14 <8.58 <0.97 0.22±0.14 ... <7.72 <0.24 <8.04 <0.42 -0.03±0.14 <8.04 <0.24 -0.01±0.14 <8.45 <0.82 0.02±0.14 <9.26 <1.89 0.06±0.14 0.26 0.15±0.20 7.86 6.68 -0.05 0.08±0.20 <8.80 <1.36 0.46±0.14 ... <8.36 <0.89 <8.36 <0.54 0.16±0.14 8.49 0.87 -0.05±0.14 0.23 0.00±0.20 8.11 <8.03 <0.48 -0.14±0.20 <8.09 <0.80 0.14±0.14 <8.79 <1.40 0.16±0.14 7.90 0.33 0.01±0.20 0.19 0.33 0.29 0.35 0.31 0.53 0.16 0.24 0.26 0.37 ... 0.28 0.32 0.45 0.94 0.05 0.43 0.73 ... 0.37 0.48 0.01 0.40 0.53 0.70 0.35 -0.33 <1.44 0.34 <1.26 0.58 <1.26 <0.63 0.37 <0.68 <0.81 0.36 1.10 <0.98 <1.11 ... <0.18 -0.14 <1.12 <1.18 <1.00 <1.00 <0.16 -0.06 1.65 1.67 0.17 The Li I lines at λ6708 Å in the present sample are very weak. For 14 of them it was possible to derive an upper limit for Li abundances. Table 13 shows the Li abundances and Figure 4 shows their behaviour with temperature. The abundances usu- ally obtained for dwarfs are higher than for giants, even taking into account that in several cases it was possible to compute only an upper limit. For the dwarf star BD+18 5215, Te f f = 6300 K and log ǫ(Li) = 1.67 and for the coolest star in the sam- ple HD 123396, Te f f = 4360 K and log ǫ(Li) = -0.14. 4.2. Carbon 4.3. Nitrogen C abundances were derived using molecular synthesis of the C2 Swan (0,0) λ5165.2 Å, C2 Swan (0,1) λ5635.5 Å band heads and the G band (CH A3∆ - X3π) at λ4290-4315 Å. Examples of spectrum synthesis are shown in Figure 7. Figure 5 shows that the sample stars are C-rich as compared with normal stars of the disk and halo, this being characteris- tic of barium stars (Bidelman & Keenan 1951; Warner 1965). Despite the dispersion, the results show an increasing trend in [C/Fe] toward lower metallicities, except for the halo giant HD 123396. Barbuy et al. (1992) derived abundances of C, N and O for a sample of barium stars, including the star HD 27271 in com- mon with the present sample. Through the synthesis of the C2 line at λ5136 Å, they found [C/Fe] = 0.15, compatible with the present work, [C/Fe] = 0.23, that is an average of several lines. They obtained the range -0.25 ≤ [C/Fe] ≤ 0.3 taking into account all stars in their sample and all indicators, and only 3 stars could be deficient in C. For the present sample the range -0.02 ≤ [C/Fe] ≤ 0.79 was found, with [C/Fe] < 0 for one star. Pereira (2005) determined C abundances for the stars HD 8270, Reddy et al. (2003) derived N abundances from two N I lines for F and G dwarfs, and observed a large dispersion -0.05 ≤ [N/Fe] ≤ 0.6 at -0.4 ≤ [Fe/H] ≤ 0.15. The sample of Clegg et al. (1981) resulted in [N/Fe] ≈ 0 in the same range of metallicities. In the present work, nitrogen abundances were determined by synthesis of CN lines in the regions λλ6476 - 6480 Å, λλ6703 - 6709 Å and λλ8030 - 8041 Å, with C abundance pre- viously established from the average of the synthesis of C2 and CH (Sect. 4.2). Figure 8 illustrates the fits to CN bands. The CN bands are very weak in stars with higher temperatures, and for them only an upper limit was given, as indicated in Figure 5 and Table 13. The results are shown in Tables 12 and 13. [N/Fe] seems to increase toward lower metallicities, except for the star HD 123396, as shown in Figure 5. The dispersion is probably due to different degrees of mixing. Barbuy et al. (1992) noted that the barium giants of their sample are rich in both N and C. Figure 5 and Table 13 show that the same applies to the present sample. For the star in common HD 27271, a nitrogen abundance 0.23 dex lower than D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 13 Barbuy et al. (1992) was found. Figure 5 suggests that the less evolved barium stars show larger N overabundances. The CNO excess provides clues on the origin of the ma- terial rich in heavy elements that polluted the envelope of the barium stars, by comparing them with normal stars. During the main sequence, the H burning through CNO cycle conserves the number of nuclei of C, N and O. During the first dredge-up the atoms of 12C capture protons forming 13C or 14N, and, as a result, the 12C is depleted by ≈ 0.2 dex and the N increases by ≈ 0.3 dex. Some normal giants can reach [N/Fe] = 0.55 and in metal-poor stars the mixing is more efficient. The barium gi- ants have already passed by the first dredge-up, and therefore they should carry the characteristics of this event in addition to their peculiarities. Once the O abundance is not modified and N increases for every star after the first dredge-up, the C is consid- ered the main responsible for the CNO excess in barium stars relative to normal stars, suggesting that the origin of the mate- rial that has polluted the envelope is the He burning shell. The fact that the less evolved barium stars are rich in N suggests that N is also responsible for the CNO excess in these stars. Table log g. The range found by Barbuy et al. (1992) for their giants 13 shows [PCNO/Fe] > 0 for all barium stars, and Figure 6 shows a clear increasing trend of [PCNO/Fe] with Te f f and correspond to 0.06 ≤ [PCNO/Fe] ≤ 0.24 and for the present work, 0.01 ≤ [PCNO/Fe] ≤ 0.73, with the less evolved stars showing the higher values. 4.4. SodiumandAluminum It is probable that the production of Na, Mg and Al occurs through C and N burning in massive stars, and for this reason, probably the SN II are the main sources of α-, Na and Al in the disk. Sharing the same production site, the pattern of abun- dances are expected to show similarities, and Figure 11 shows that it is true for the program stars. For the same reason, it is in- teresting to study the relations involving [Al/Mg] and [Na/Mg] in addition to [Al/Fe] and [Na/Fe]. The Baumuller & Gehren (1997) analysis taking into ac- count NLTE effects suggests that for [Fe/H] ≈ -0.5 there is an overabundance of [Al/Fe] ≈ 0.15 dex. According to Figure 11 and Table 14, some of the sample stars are found in this re- gion, one of them being the halo star HD 5424 ([Fe/H] = -0.55). The others are below this value, except HD 210910 ([Fe/H] = -0.37), which is a subgiant with broad lines, and HD 123396 ([Fe/H] = -1.19), which is a halo giant. Most data are found in the range -0.1 ≤ [Al/Fe] ≤ 0.1. Relative to Mg, Figure 9 shows that in the range -0.75 ≤ [Mg/H] ≤ 0, 4 stars show overabun- dance and 19 a deficiency of Al relative to Mg, with values in the range -0.4 ≤ [Al/Mg] ≤ +0.1. In the Al abundance calcula- tion the lines λ6696 Å and λ6698.7 Å were used. Baumuller et al. (1998) analysed a sample of stars taking into account NLTE effects and verified a decreasing trend of Na abundance toward decreasing metallicities. The present sample stars are found to be in the range -0.18 ≤ [Na/Fe] ≤ 0.31, show- ing no clear trend, but most data have -0.1 ≤ [Na/Fe] ≤ 0.2, as shown in Figure 11. Relative to Mg, Figure 9 shows that the results are in the range -0.4 ≤ [Na/Mg] ≤ +0.35 with a larger dispersion than [Al/Mg]. A Na deficiency relative to Mg was found in 11 stars, and for the other stars [Na/Mg] ≥ 0. Six lines of Na I with two doublets were used. For 7 stars the difference between the lines is within ± 0.25 and ± 0.30. For one star, the difference is 0.40 dex. For the other sample stars there is good agreement between the lines. The Al and Na excesses relative to Ba show an increase toward higher metallicities (see Figure 16 and Table 18), while a decreasing trend is seen relative to [Ba/H] (Figure 15). In the range 0 ≤ [Ba/H] ≤ 1.5, -0.2 ≤ [Na/Ba] ≤ -1.6 and -0.3 ≤ [Al/Ba] ≤ -1.8 were found. Otherwise, the excesses of Al and Na relative to Eu show no trend as a function of [Fe/H], with -0.7 ≤ [Na/Eu] ≤ 0.2 and -0.8 ≤ [Al/Eu] ≤ 0 (Figure 18). In the range -0.7 ≤ [Eu/H] ≤ 0.4 there is a small decreasing trend of [Al,Na/Eu] toward higher [Eu/H] (see Figure 17 and Table 18). Table 15. Lines, equivalent widths and abundances. Symbols: '+': abundances derived by using metallicities from Fe  lines; '*': multiple line; '**': Ni line blended with O line. < in- dicates an upper limit for oxygen. The gf-values sources are: 1 - Allende Prieto et al. (2001), 2 - Lambert (1978), 3 - NIST, 4 - Fuhrmann et al. (1995), 5 - Prochaska et al. (2000), 6 - McWilliam & Rich (1994), 7 - Barbuy et al. (1999), 8 - Bielski (1975), 9 - Bi´emont & Godefroid (1980), 10 - Gratton & Sneden (1994), 11 - Hannaford et al. (1982), 12 - Hannaford & Lowe (1983), 13 - Bi´emont et al. (1981), 14 - Th´evenin (1990), 15 - Th´evenin (1989), 16 - Smith et al. (2000), 17 - McWilliam (1998), 18 - Rutten (1978), 19 - Lawler et al. (2001a), 20 - Palmeri et al. (2000), 21 - Goly et al. (1991), 22 - Lage & Whaling (1976), 23 - Hartog et al. (2003), 24 - Maier & Whaling (1977), 25 - Sneden et al. (1996), 26 - Bi´emont et al. (1989), 27 - Lawler et al. (2001b), 28 - Bergstrom et al. (1988), 29 - Corliss & Bozman (1962), 30 - mean between Kusz (1992) and Bi´emont & Lowe (1993), 31 - Bi´emont et al. (2000) 32 - Bi´emont et al. (1983). Full table is only available in electronic form. HD 749 HR 107 λ el 6300.31 O I O I 6363.776 Na I 4982.808 Na I 4982.813 Na I 5682.65 Na I 5688.193 . . . . χex 0.00 0.02 2.10 2.10 2.10 2.104 . . log gf -9.717 -10.250 -1.913 -0.961 -0.700 -1.390 . . ref 1 2 3 3 3 3 . . EW log ǫ 27 8.94 8.84 17 6.03 102 6.03 * 6.08 104 135 6.18 . . . . [X/Fe] 0.03 -0.07 -0.47 -0.47 -0.42 -0.32 . . [X/Fe]+ 0.26 0.16 -0.24 -0.24 -0.19 -0.09 . . EW log ǫ 6 8.31 ... ... 5.97 39 5.97 * 5.94 55 86 6.04 . . . . [X/Fe] -0.07 ... ... 0.00 ... 0.00 ... -0.03 ... 0.07 ... . . 4.5. α-elements The observations show that the α-elements tend to be overabun- dant at low metallicities, with [X/Fe] reaching ≈ 0.5 at -4 < [Fe/H] < -1 (e.g. Barbuy 1988; Cayrel et al. 2004). At [Fe/H] ≈ -1 the overabundance starts to decrease toward higher metallic- ities, and [X/Fe] can be subsolar at [Fe/H] ≈ 0. This behaviour of the α-elements relative to iron has been attributed to the time delay between core-collapse supernovae and SNIa. 14 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Fig. 9. [Al/Mg] and [Na/Mg] vs. [Mg/H]. Symbols are the same as in Figure 4. Fig. 10. [α,pFe/Fe] vs. [Fe/H]. Symbols are the same as in Figure 4. No substantial difference between α-elements in normal and barium stars were seen up to now. Figure 10 shows [α/Fe] vs. [Fe/H] for the sample barium stars where α's included are O, Mg, Si, Ca and Ti. The solar abundance adopted was log ǫ(O) = 8.74, which is the value by Asplund et al. (2005) through a 3D atmosphere model, log ǫ(O) = 8.66, corrected by 0.08 for 1D, according to Allende Prieto et al. (2001). Oxygen: It is the third most abundant element in the Universe after H and He. It is produced during He burning in the interior of massive stars, and is released through SN II events. Abundances of oxygen for the present sample were deter- mined through spectrum synthesis of the forbidden lines of [O I] at λ6300.3 Å and λ6363.8 Å. These lines are reliable since they are not subject to NLTE effects. The resolution of the pro- gram stars spectra is such that the Sc  line at λ6300.7 Å is not blended with the [O I] line at λ6300.3 Å. Telluric lines are displaced thanks to their radial velocities and do not blend the oxygen line in the sample stars. For 2 stars, HD 147609 and HD 87080, the sky emission line is present preventing the de- termination of the O abundance. For HD 13551, HD 22589 and HD 107574, an upper limit was derived. The Ni I line at λ6300.335 Å was taken into account in the oxygen abundance calculations. The adopted abundance for Ni was the average of other ten lines and the atomic constants are shown in Table 15. The O abundance results obtained shown in Figure 11, are in good agreement with Figure 4 by Franc¸ois et al. (2004) in the same range of metallicities. Magnesium: Similarly to oxygen, magnesium is also pro- duced by massive stars, but in this case carbon and neon burn- ing is responsible for its production. Figure 4 of Franc¸ois et al. (2004) shows the evolution of [Mg/Fe] relative to [Fe/H], where [Mg/Fe] decreases toward increasing metallicities. The 3 lines used in the determination of Mg abundance for the present sample present a good agreement, as shown in Table 15. Figure 11 shows that the range of [Mg/Fe] is simi- lar to those of [Na/Fe] and [Al/Fe], differing from [Ti/Fe], that reaches lower abundances, in agreement with Franc¸ois et al. (2004, and references therein). Silicon: SNae of ≈ 20 M could be the main sources of Si (Woosley & Weaver 1995). Prochaska et al. (2000) found a high overabundance of Si and an increasing abundance trend toward lower metallicities for stars of the thick disk in the ⊙ Table 14. Abundance ratios of Al and Na relative to Mg, and the α and pFe relative to Fe. D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 15 star [Al/Mg] σ[Al/Mg] [Na/Mg] σ[Na/Mg] [Mg/H] σ[Mg/H] [α/Fe] σ[α/Fe] [pFe/Fe] σ[pFe/Fe] HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 HD 106191 HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 HD 210709 HD 210910 HD 222349 BD+18 5215 HD 223938 0.00 -0.10 -0.05 -0.01 -0.05 -0.20 -0.11 0.03 -0.20 -0.29 -0.11 -0.37 -0.14 -0.09 0.00 0.04 -0.21 0.01 -0.05 0.00 -0.11 -0.06 0.06 -0.18 -0.27 -0.17 0.20 0.10 0.20 0.10 0.20 0.10 0.10 0.20 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.20 0.20 0.10 0.10 0.10 0.10 0.20 0.20 0.10 0.10 0.20 -0.06 -0.05 -0.10 0.09 0.19 -0.14 0.12 0.34 -0.05 0.09 -0.05 -0.12 0.12 0.01 0.07 0.02 -0.34 0.13 0.14 0.17 0.15 -0.10 -0.28 0.03 0.00 -0.05 0.14 0.07 0.14 0.07 0.14 0.07 0.07 0.14 0.07 0.07 0.07 0.07 0.07 0.07 0.07 0.14 0.14 0.07 0.07 0.07 0.07 0.14 0.14 0.07 0.07 0.14 -0.18 -0.28 -0.35 -0.46 -0.07 -0.19 -0.06 -0.21 -0.60 -0.32 -0.40 -0.23 -0.20 -0.22 -0.47 -0.36 -0.73 -0.44 -0.49 -0.25 -0.31 0.00 -0.26 -0.50 -0.36 -0.26 0.21 0.07 0.21 0.07 0.21 0.07 0.07 0.21 0.07 0.07 0.07 0.07 0.07 0.07 0.07 0.21 0.21 0.07 0.07 0.07 0.07 0.21 0.21 0.07 0.07 0.21 0.17 -0.08 0.16 0.04 0.13 0.38 0.02 0.14 0.00 0.09 -0.01 0.23 0.27 0.32 0.28 -0.07 0.39 0.27 0.01 0.22 0.38 -0.05 0.39 0.37 0.38 0.30 0.28 0.11 0.28 0.11 0.28 0.11 0.11 0.28 0.11 0.11 0.11 0.11 0.11 0.11 0.11 0.28 0.28 0.11 0.11 0.11 0.11 0.28 0.28 0.11 0.11 0.28 0.06 -0.04 0.09 -0.07 0.12 0.00 0.15 -0.01 -0.14 -0.04 -0.07 -0.03 -0.09 0.07 -0.08 -0.09 0.02 -0.02 -0.02 -0.06 0.00 -0.03 -0.09 -0.04 0.03 -0.02 0.24 0.06 0.24 0.06 0.24 0.06 0.06 0.24 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0.24 0.24 0.06 0.06 0.06 0.06 0.24 0.24 0.06 0.06 0.24 range of metallicities -1.2 < [Fe/H] < -0.3, in agreement with McWilliam (1997). For the present sample 5 lines were used. All lines give sim- ilar abundances, except the line λ5948.5 Å that gives values 0.5 dex higher for some stars, possibly due to an unknown blend. Most results give [Si/Fe] ≈ 0, except for the star HD 123396, a halo giant, for which the abundance is higher, as shown in Figure 11. The results are in agreement with the chemical evo- lution model by Franc¸ois et al. (2004) for this range of metal- licities. The odd-even effect observed by Arnett (1971), where Mg and Si are overabundant relative to Na and Al, is confirmed for some sample stars. For several stars, Na and Al abundances are higher than Si and Mg, with -0.1 ≤ [Mg/Fe] ≤ 0.2, -0.15 ≤ [Si/Fe] ≤ 0.2, -0.1 ≤ [Na/Fe] ≤ 0.2 and -0.1 ≤ [Al/Fe] ≤ 0.1. Calcium: According to McWilliam (1997), the abundance of Ca is expected to behave similarly to Si, given that SN II of moderate mass is the main source of both elements (Woosley & Weaver 1995), and they can be also released by SN Ia (Nomoto et al. 1997b). The similarity of behaviour be- tween Ca and Si was verified by Prochaska et al. (2000) as well as for the barium stars of the present sample (Figure 11), except for the star HD 210910. Six Ca lines were used, being some of them very strong in some sample stars (Table 15). For 16 stars the difference between abundances derived from those lines is from 0.2 to 0.5 dex, and for the others, there is a better agree- ment. Titanium: Ti is usually included in the α-elements list be- cause its overabundance in metal-poor stars is similar to that of α-elements (Gratton & Sneden 1991), but its nucleosynthesis is unclear (Woosley & Weaver 1995). For this reason, Ti abun- dance can be different from that of Ca and Si, as observed by Prochaska et al. (2000). According to Franc¸ois et al. (2004), Ti and Mg abundances have a similar behaviour, though the dis- persion is larger for Mg in their Figure 4. In the present work, Ti abundances are usually lower than those of Ca, Si and Mg, as shown in Figures 11 and 12. Twelve lines of Ti I were used. For 4 stars, the abundance from the λ5210.4 Å line is different from the other lines. The reason for that difference is unclear. If the problem were in the atomic constants, the difference would be observed for all stars. In gen- eral, the abundance results from the 12 lines are in agreement. 4.6. Ironpeakelements At the last moments of the life of a massive star, the iron and the iron peak elements are formed in large amounts (Woosley & Weaver 1995). The process that precedes the ex- plosion of SN Ia also produces these elements, but in lower amounts relative to massive stars. However, the SN Ia ejecta contain larger amounts of those elements than SN II, because part of the yield is restrained in the neutron star newly formed. Figure 12 of McWilliam (1997) shows abundances from previous work, and it is possible to observe the trends of [X/Fe] 16 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Table 16. Mean logǫ(X) and [X/Fe]. The symbol "*" indicates that the metallicity was derived from Fe  lines. HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 el log ǫ(X) [X/Fe] [X/Fe]* log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] [X/Fe]* log ǫ(X) [X/Fe] O Na Mg Al Si Ca Sc Ti V Cr Co Ni Cu Zn Sr  Sr  Y Zr  Zr  Mo Ba La Ce Pr Nd Sm Eu Gd Dy Pb 8.89 6.09 7.40 6.29 7.66 6.28 3.42 4.75 3.93 5.51 5.37 6.25 4.20 4.53 3.49 3.59 3.36 2.86 3.99 1.99 3.25 2.33 3.14 1.49 2.72 1.88 0.79 1.38 2.06 2.27 -0.02 -0.41 -0.35 -0.35 -0.06 -0.25 0.08 -0.44 -0.24 -0.33 0.28 -0.17 -0.18 -0.24 0.35 0.45 0.95 0.09 1.22 -0.10 0.95 1.03 1.27 0.66 1.10 0.70 0.10 0.09 0.69 0.15 0.21 -0.18 -0.12 -0.12 0.17 -0.02 0.31 -0.21 -0.01 -0.10 0.51 0.06 0.05 -0.01 0.58 0.68 1.18 0.32 1.45 0.13 1.18 1.26 1.50 0.89 1.33 0.93 0.33 0.32 0.92 0.38 8.31 6.00 7.30 6.09 7.13 6.04 3.00 4.69 3.70 5.28 4.74 5.83 3.68 4.14 2.91 3.36 2.48 2.84 2.81 2.16 2.72 1.44 1.75 0.82 1.41 0.95 0.20 1.31 ... 2.49 -0.07 0.03 0.08 -0.02 -0.06 0.04 0.19 0.03 0.06 -0.03 0.18 -0.06 -0.17 -0.10 0.30 0.75 0.60 0.60 0.57 0.60 0.95 0.67 0.41 0.52 0.32 0.30 0.04 0.55 ... 0.90 8.37 5.88 7.23 6.07 7.20 5.71 3.11 4.41 3.28 5.14 4.66 5.79 3.49 4.15 3.37 3.02 2.72 2.54 3.40 1.57 3.06 2.13 3.01 1.49 2.62 1.73 0.43 1.02 2.30 2.50 0.18 0.10 0.20 0.15 0.20 -0.10 0.49 -0.06 -0.17 0.02 0.29 0.09 -0.17 0.10 0.95 0.60 1.03 0.49 1.35 0.20 1.48 1.55 1.86 1.38 1.72 1.27 0.46 0.45 1.65 1.10 8.40 5.96 7.12 6.00 7.11 6.01 3.01 4.47 3.36 5.13 4.43 5.77 3.58 4.14 3.38 3.45 2.77 3.02 3.18 2.10 2.82 1.71 2.11 0.72 1.76 0.96 0.42 0.95 0.82 2.03 0.08 0.05 -0.04 -0.05 -0.02 0.07 0.26 -0.13 -0.22 -0.12 -0.07 -0.06 -0.21 -0.04 0.83 0.90 0.95 0.84 1.00 0.60 1.11 1.00 0.83 0.48 0.73 0.37 0.32 0.25 0.04 0.50 8.80 6.45 7.51 6.35 7.34 6.23 3.48 5.03 3.94 5.64 5.27 6.23 4.18 4.50 3.90 3.63 3.33 3.20 3.84 2.30 3.52 2.62 3.25 1.93 2.82 2.36 0.88 1.57 1.83 2.98 0.18 0.24 0.05 0.00 -0.09 -0.01 0.43 0.13 0.06 0.09 0.47 0.10 0.09 0.02 1.05 0.78 1.21 0.72 1.36 0.50 1.51 1.61 1.67 1.39 1.49 1.47 0.48 0.57 0.75 1.15 <8.73 <0.23 -0.09 0.05 -0.15 -0.10 0.00 0.05 -0.25 -0.18 -0.26 -0.20 -0.19 -0.37 -0.10 0.56 0.80 0.88 0.69 0.82 0.60 0.96 0.79 0.71 0.37 0.53 0.25 0.01 0.35 -0.11 0.30 6.00 7.39 6.08 7.21 6.12 2.98 4.53 3.58 5.17 4.48 5.82 3.60 4.26 3.29 3.53 2.88 3.05 3.18 2.28 2.85 1.68 2.17 0.79 1.74 1.02 0.29 1.23 0.85 2.01 <0.43 0.11 0.25 0.05 0.10 0.20 0.25 -0.05 0.02 -0.06 0.00 0.01 -0.17 0.10 0.76 1.00 1.08 0.89 1.02 0.80 1.16 0.99 0.91 0.57 0.73 0.45 0.21 0.55 0.09 0.50 <8.50 <0.03 0.33 0.21 0.10 0.04 0.35 0.16 0.02 -0.08 0.23 0.08 0.13 0.00 0.11 0.83 0.88 0.83 0.49 1.07 0.20 0.88 0.70 0.45 0.27 0.32 0.08 0.21 0.03 0.04 -0.15 6.39 7.52 6.30 7.32 6.44 3.06 4.77 3.65 5.63 4.73 6.11 3.94 4.44 3.53 3.58 2.80 2.82 3.40 1.85 2.74 1.56 1.88 0.66 1.50 0.82 0.46 0.88 0.97 1.53 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 el log ǫ(X) [X/Fe] [X/Fe]* log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] HD 106191 log ǫ(X) [X/Fe] O Na Mg Al Si Ca Sc Ti V Cr Co Ni Cu Zn Sr  Sr  Y Zr  Zr  Mo Ba La Ce Pr Nd Sm Eu Gd Dy Pb 8.84 6.46 7.37 6.29 7.45 6.20 3.42 4.69 3.63 5.35 5.00 6.19 4.01 4.47 3.37 3.52 3.04 2.65 3.55 1.79 2.92 1.76 2.27 1.06 1.89 1.33 0.74 1.28 1.31 2.17 -0.07 -0.04 -0.38 -0.35 -0.27 -0.33 0.08 -0.50 -0.54 -0.49 -0.09 -0.23 -0.37 -0.30 0.23 0.38 0.63 -0.12 0.78 -0.30 0.62 0.46 0.40 0.23 0.27 0.15 0.05 -0.01 -0.06 0.05 0.19 0.22 -0.12 -0.09 -0.01 -0.07 0.34 -0.24 -0.28 -0.23 0.17 0.03 -0.11 -0.04 0.49 0.64 0.89 0.14 1.04 -0.04 0.88 0.72 0.66 0.49 0.53 0.41 0.31 0.25 0.20 0.31 8.15 5.68 6.98 5.67 6.87 5.79 2.82 4.35 3.19 4.88 4.22 5.50 3.27 4.01 3.32 3.39 2.63 2.71 3.17 1.80 2.80 1.90 2.68 1.13 2.14 1.34 0.25 1.19 1.27 2.68 0.03 -0.03 0.02 -0.18 -0.06 0.05 0.27 -0.05 -0.19 -0.17 -0.08 -0.13 -0.32 0.03 0.97 1.04 1.01 0.73 1.19 0.50 1.29 1.39 1.60 1.09 1.31 0.95 0.35 0.69 0.69 1.35 8.56 6.10 7.26 5.86 7.21 6.16 3.09 4.59 3.46 5.24 4.64 5.91 3.78 4.27 3.74 3.92 3.10 3.47 3.55 2.21 3.17 2.04 2.45 1.10 1.91 1.23 0.46 1.25 1.18 2.49 0.13 0.08 -0.01 -0.30 -0.03 0.11 0.23 -0.12 -0.23 -0.12 0.03 -0.03 -0.12 -0.02 1.08 1.26 1.17 1.18 1.26 0.60 1.35 1.22 1.06 0.75 0.77 0.53 0.25 0.44 0.29 0.85 ... 5.88 7.18 5.96 7.02 6.04 3.13 4.53 3.42 5.13 4.48 5.74 3.52 4.26 3.51 3.53 2.91 2.96 3.51 1.88 3.17 2.43 2.99 1.46 2.57 1.69 0.74 1.58 1.90 2.56 ... -0.01 0.04 -0.07 -0.09 0.12 0.40 -0.05 -0.14 -0.10 0.00 -0.07 -0.25 0.10 0.98 1.00 1.11 0.80 1.35 0.40 1.48 1.74 1.73 1.24 1.56 1.12 0.66 0.90 1.14 1.05 8.72 5.98 7.35 5.87 7.19 6.10 2.95 4.56 3.60 5.37 4.60 5.91 3.74 4.32 3.65 3.75 2.96 3.15 3.33 2.12 2.82 1.76 2.11 0.96 1.80 1.14 0.38 1.07 0.59 2.00 0.28 -0.05 0.07 -0.30 -0.06 0.04 0.08 -0.16 -0.10 0.00 -0.02 -0.04 -0.17 0.02 0.98 1.08 1.02 0.85 1.03 0.50 0.99 0.93 0.71 0.60 0.65 0.43 0.16 0.25 -0.31 0.35 8.95 6.25 7.38 6.13 7.29 6.32 3.23 4.78 3.77 5.42 4.82 6.04 3.82 4.33 3.52 3.58 2.76 3.03 3.23 2.20 3.05 1.73 2.18 0.98 1.75 1.13 0.72 1.14 1.17 2.53 0.33 0.04 -0.08 -0.22 -0.14 0.08 0.18 -0.12 -0.11 -0.13 0.02 -0.09 -0.27 -0.15 0.67 0.73 0.64 0.55 0.75 0.40 1.04 0.72 0.60 0.44 0.42 0.24 0.32 0.14 0.09 0.70 8.83 6.12 7.36 6.16 7.18 6.17 3.03 4.70 3.77 5.45 4.78 6.03 3.88 4.39 3.63 3.33 2.86 ... 3.38 2.13 2.72 1.50 2.02 1.21 1.64 1.18 0.43 1.23 1.20 2.31 0.38 0.08 0.07 -0.02 -0.08 0.10 0.15 -0.03 0.06 0.07 0.15 0.07 -0.04 0.08 0.95 0.65 0.91 ... 1.07 0.50 0.88 0.66 0.61 0.84 0.48 0.46 0.20 0.40 0.29 0.65 The usual notations were adopted: log ǫ(A) = log(NA/NH)+12 and [A/B] = log(NA/NB) particle density of "A" and "B", respectively. ∗ -log(NA/NB) ⊙ , where NA and NB are the numerical relative to metallicity for iron peak elements. In the present work, this trend is unclear. However, in average, [pFe/Fe] ≈ 0 at -1.2 < [Fe/H] < 0, as shown in Figure 10, being pFe the average of V, Cr, Co and Ni abundances. The stellar yield of iron peak elements is uncertain given that it depends on several process not well established such as the amount of mass released during supernovae events, the mass retained in the proto neutron star, the energy of the explo- sion and the neutron flux. D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 17 Table 17. Same as Table 16 for other 12 sample stars and log ǫ(X) for the Sun. For solar abundances the errors on last decimals are given in parenthesis. HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 el log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] [X/Fe]* log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] [X/Fe]* log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] O Na Mg Al Si Ca Sc Ti V Cr Co Ni Cu Zn Sr  Sr  Y Zr  Zr  Mo Ba La Ce Pr Nd Sm Eu Gd Dy Pb <8.52 <0.33 0.15 0.08 0.08 0.01 0.07 0.08 -0.04 0.11 -0.15 0.07 -0.07 -0.22 -0.04 0.60 1.10 0.96 ... 0.95 0.90 1.71 1.16 1.02 0.70 0.86 0.64 0.47 0.36 0.34 1.05 5.93 7.11 6.00 7.01 5.88 2.70 4.43 3.56 4.97 4.44 5.63 3.44 4.01 3.02 3.52 2.65 ... 3.00 2.27 3.29 1.74 2.17 0.81 1.76 1.10 0.44 0.93 0.99 2.45 8.38 5.99 7.22 6.15 7.21 5.88 2.98 4.54 3.41 5.12 4.73 5.86 3.65 4.25 3.08 2.90 2.51 2.31 2.96 1.40 2.83 1.77 2.28 0.98 1.99 1.25 0.36 1.02 1.43 2.48 -0.04 -0.02 -0.04 0.00 -0.02 -0.16 0.13 -0.16 -0.27 -0.23 0.13 -0.07 -0.24 -0.03 0.43 0.25 0.59 0.03 0.68 -0.20 1.02 0.96 0.90 0.64 0.86 0.56 0.16 0.22 0.55 0.85 7.97 5.26 6.85 5.53 6.65 5.25 2.45 3.76 2.48 4.25 3.90 5.12 2.83 3.84 2.40 2.18 1.75 1.31 2.60 0.53 2.24 1.21 2.13 0.67 1.81 1.01 -0.17 0.74 1.20 1.96 0.22 -0.08 0.26 0.05 0.09 -0.12 0.27 -0.27 -0.53 -0.43 -0.03 -0.14 -0.39 0.23 0.42 0.20 0.50 -0.30 0.99 -0.40 1.10 1.07 1.42 1.00 1.35 0.99 0.30 0.61 0.99 1.00 0.42 0.12 0.46 0.25 0.29 0.08 0.47 -0.07 -0.33 -0.23 0.17 0.06 -0.19 0.43 0.62 0.40 0.70 -0.10 1.19 -0.20 1.30 1.27 1.62 1.20 1.55 1.19 0.50 0.81 1.19 1.20 8.59 6.02 7.14 6.04 7.01 5.95 2.89 4.59 3.63 5.12 4.78 5.74 3.77 4.22 3.59 3.70 3.10 3.57 3.48 2.44 3.44 2.30 2.91 1.49 2.38 1.73 0.87 1.19 1.36 3.02 0.33 0.17 0.04 0.05 -0.06 0.07 0.20 0.05 0.11 -0.07 0.34 -0.03 0.04 0.10 1.10 1.21 1.34 1.45 1.36 1.00 1.79 1.65 1.69 1.31 1.41 1.20 0.83 0.55 0.64 1.55 ... 5.98 7.09 5.93 7.16 5.88 3.43 4.41 3.35 5.12 4.46 5.80 3.66 4.29 3.56 4.11 3.36 3.04 3.71 2.20 3.25 2.31 2.89 1.43 2.32 1.65 0.81 1.47 1.45 2.28 ... -0.43 -0.57 -0.62 -0.47 -0.56 0.18 -0.69 -0.73 -0.63 -0.54 -0.53 -0.63 -0.39 0.51 1.06 1.04 0.36 1.03 0.20 1.04 1.10 1.11 0.69 0.79 0.56 0.21 0.27 0.17 0.25 ... 0.10 -0.04 -0.09 0.06 -0.03 0.71 -0.16 -0.20 -0.10 -0.01 0.00 -0.10 0.14 1.04 1.59 1.57 0.89 1.56 0.73 1.57 1.63 1.64 1.22 1.32 1.09 0.74 0.80 0.70 0.78 8.92 6.25 7.33 6.22 7.32 6.21 3.32 4.86 3.78 5.50 4.90 6.09 3.99 4.50 3.68 3.57 3.22 3.51 3.62 2.32 3.06 1.83 2.15 1.01 1.69 1.14 0.62 1.25 1.24 2.55 0.28 0.02 -0.15 -0.15 -0.13 -0.05 0.25 -0.06 -0.12 -0.07 0.08 -0.06 -0.12 0.00 0.81 0.70 1.08 1.01 1.12 0.50 1.03 0.80 0.55 0.45 0.34 0.23 0.20 0.23 0.14 0.70 8.88 6.17 7.27 6.05 7.19 6.14 3.11 4.70 3.75 5.41 4.68 5.93 3.80 4.37 3.72 3.65 2.96 2.95 3.48 2.12 3.03 1.99 2.63 1.18 2.19 1.43 0.51 1.36 1.09 2.60 0.44 0.14 -0.01 -0.12 -0.06 0.08 0.24 -0.02 0.05 0.04 0.06 -0.02 -0.11 0.07 1.05 0.98 1.02 0.65 1.18 0.50 1.20 1.16 1.23 0.82 1.04 0.72 0.29 0.54 0.19 0.95 HD 210709 HD 210910 HD 222349 BD+18 5215 HD 223938 Sun el log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] [X/Fe]* log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] log ǫ(X) [X/Fe] [X/Fe]* log ǫ(X) ref O Na Mg Al Si Ca Sc Ti V Cr Co Ni Cu Zn Sr  Sr  Y Zr  Zr  Mo Ba La Ce Pr Nd Sm Eu Gd Dy Pb 8.68 6.23 7.58 6.41 7.39 6.11 3.31 4.73 3.92 5.60 5.21 6.15 4.00 4.40 3.26 3.02 2.73 2.47 3.35 1.58 2.82 1.77 2.40 1.01 2.04 1.27 0.56 1.03 1.31 2.36 -0.02 -0.06 0.04 -0.02 -0.12 -0.21 0.18 -0.25 -0.04 -0.03 0.33 -0.06 -0.17 -0.16 0.33 0.09 0.53 -0.09 0.79 -0.30 0.73 0.68 0.74 0.39 0.63 0.30 0.08 -0.05 0.15 0.45 8.81 5.79 7.32 6.27 7.35 5.54 3.08 4.14 3.35 4.62 4.69 5.86 3.68 4.41 3.06 3.46 2.41 2.12 2.59 1.36 2.75 1.50 1.98 1.15 1.63 1.05 0.69 1.31 1.13 1.54 0.03 -0.58 -0.30 -0.24 -0.24 -0.86 -0.13 -0.92 -0.69 -1.09 -0.27 -0.43 -0.57 -0.23 0.05 0.45 0.13 -0.52 -0.05 -0.60 0.58 0.33 0.24 0.45 0.14 0.00 0.13 0.15 -0.11 -0.45 0.44 -0.17 0.11 0.17 0.17 -0.45 0.28 -0.51 -0.28 -0.68 0.14 -0.02 -0.16 0.18 0.46 0.86 0.54 -0.11 0.36 -0.19 0.99 0.74 0.65 0.86 0.55 0.41 0.54 0.56 0.30 -0.04 8.54 5.86 7.08 5.79 6.89 5.82 2.76 4.34 3.51 4.90 4.26 5.60 3.39 4.01 3.24 3.39 2.64 3.20 3.19 1.79 2.88 1.85 2.47 0.92 2.08 1.26 0.13 1.06 1.16 2.77 0.43 0.16 0.13 -0.05 -0.03 0.09 0.22 -0.05 0.14 -0.14 -0.03 -0.02 -0.19 0.04 0.90 1.05 1.03 1.23 1.22 0.50 1.38 1.35 1.40 0.89 1.26 0.88 0.24 0.57 0.59 1.45 8.64 5.97 7.22 5.84 7.07 5.90 2.82 4.45 3.56 5.14 4.74 5.74 3.59 4.12 3.55 3.59 2.72 3.37 3.32 2.09 3.06 1.79 2.28 1.01 1.75 1.18 0.23 1.74 1.21 1.87 0.43 0.17 0.17 -0.10 0.05 0.07 0.18 -0.04 0.09 0.00 0.35 0.02 -0.09 0.05 1.11 1.15 1.01 1.30 1.25 0.70 1.46 1.19 1.11 0.88 0.83 0.70 0.24 1.15 0.54 0.45 8.74 6.02 7.32 6.04 7.23 6.07 3.14 4.51 3.46 5.14 4.70 5.91 3.71 4.32 3.21 3.23 2.63 2.41 3.31 1.49 3.00 1.82 2.39 0.88 2.18 1.41 0.54 1.14 1.36 2.57 0.13 -0.18 -0.13 -0.30 -0.19 -0.16 0.10 -0.38 -0.41 -0.40 -0.09 -0.21 -0.37 -0.15 0.37 0.39 0.52 -0.06 0.84 -0.30 1.00 0.82 0.82 0.35 0.86 0.53 0.15 0.15 0.29 0.75 0.35 0.04 0.09 -0.08 0.03 0.06 0.32 -0.16 -0.19 -0.18 0.13 0.01 -0.15 0.07 0.59 0.61 0.74 0.16 1.06 -0.08 1.22 1.04 1.04 0.57 1.08 0.75 0.37 0.37 0.51 0.97 8.74(6) 6.33(3) 7.58(5) 6.47(7) 7.55(5) 6.36(2) 3.17(10) 5.02(6) 4.00(2) 5.67(3) 4.92(4) 6.25(4) 4.21(4) 4.60(8) 2.97(7) 2.97(7) 2.24(3) 2.60(2) 2.60(2) 1.92(5) 2.13(5) 1.13(3) 1.70(4) 0.66(15) 1.45(1) 1.01(6) 0.52(1) 1.12(4) 1.20(6) 1.95(8) 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 3 4 5 6 7 8 2 9 2 References on solar abundances: 1 - Asplund et al. (2005); 2 - Grevesse & Sauval (1998); 3 - Lawler et al. (2001a); 4 - Palmeri et al. (2000); 5 - Lage & Whaling (1976); 6 - Hartog et al. (2003); 7 - Bi´emont et al. (1989); 8 - Lawler et al. (2001b); 9 - Bi´emont & Lowe (1993) Scandium: Zhao & Magain found Sc stars. overabundances of ≈ 0.25 dex in metal-poor Gratton & Sneden (1991) suggested that such result was (1990) have a consequence of gf-values adopted. Prochaska et al. (2000) found Sc overabundance ≈ 0.20 dex at [Fe/H] ∼ -0.5, with a decreasing trend for lower metallicities, using log gf from 18 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars [Fe/H] Fig. 11. [X/Fe] vs. [Fe/H]. The symbols carry a double information: luminosity class and ∆[Fe/H]=[FeI/H]-[FeII/H]. 4-pointed symbols (squares and stars): dwarf stars with log g ≥ 3.7; 3-pointed symbols (triangles and stars): subgiants stars with 2.4 < log g < 3.7; round symbols (circles and star): giants stars with log g ≤ 2.4. Filled symbols indicate ∆[Fe/H] < 0.2 dex, being the adopted metallicities derived from Fe  lines. Open symbols indicate the seven stars for which ∆[Fe/H] > 0.2 dex. These stars are represented twice being circles, triangles and squares corresponding to metallicities derived from Fe  lines and starred symbols, the same stars with metallicities derived from Fe  lines. The arrows in the oxygen panel indicate an upper limit for HD 13551, HD 22589 and HD 107574. Martin et al. (1988) and Lawler & Dakin (1989) for disk stars in the range -1.2 ≤ [Fe/H] ≤ -0.3. Nissen et al. (2000) found a trend of [Sc/Fe] similar that of α-elements for 100 dwarf F and G stars with -1.4 < [Fe/H] < 0, by using hyperfine structure from Steffen (1985). In the chemical evolution model of Sc by Franc¸ois et al. (2004), [Sc/Fe] ≈ 0.2 for very metal-poor stars, with a trend toward zero for stars richer than [Fe/H] ∼ -2. In the present work, 4 lines of Sc were used, with good agreement among their abundance results. The gf-values adopted were those from NIST. All results for [Sc/Fe] shown in Figure 12 are above solar, reaching 0.7 for the star HD 147609 ([Fe /H]=-0.45). Vanadium: Few studies on vanadium were found. Gratton & Sneden (1991) analysed a sample of 20 stars in an extensive range of metallicities and found [V/Fe] ∼ 0 for all D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 19 Fig. 12. [X/Fe] vs. [Fe/H]. Symbols are the same as in Figure 11. [Fe/H] metallicities, in agreement with the work by Pagel (1968). However, Prochaska et al. (2000) obtained 0.1 < [V/Fe] < 0.4. The V I lines are very sensitive to temperature. They are usually weak for all stars of the present sample, but they are weaker for hotter stars. As an example, only 2 lines are avail- able for the stars HD 123585 and BD+18 5215. Despite the strength of the lines, the abundances derived show a good agreement. A larger difference, of 0.5 dex was obtained for 2 lines of the star HD 210910 (see Table 15). All abundances are in the range -0.40 < [V/Fe] < 0.2, as shown in Figure 12. Chromium: Cayrel et al. (2004) found an increasing trend of [Cr/Fe] in the range -4 ≤ [Cr/Fe] ≤ -2 for very metal-poor stars. Figure 6 from Franc¸ois et al. (2004) shows that for higher metallicities, the data remain around [Cr/Fe] ∼ 0. In the present work, 6 lines of Cr were used, resulting in the range -0.2 ≤ [Cr/Fe] ≤ 0.2, consistent with Franc¸ois et al. (2004). For the star HD 210910 the result is lower than for other stars. Cobalt: In the same range of metallicity, Prochaska et al. (2000) found an overabundance of Co relative to Fe reaching ∼ 0.2 dex, whereas Gratton & Sneden (1991) found a deficiency of 0.1 dex. The lines of Co are sensitive to temperature. For the hotter stars of the sample, the equivalent widths are very small and in some cases, they had to be discarded. Differences in the abundances from different lines are shown in Table 15, and Figure 12 shows that -0.15 < [Co/Fe] < 0.4. Nickel: Gratton & Sneden (1991), Edvardsson et al. (1993), Peterson et al. (1990), McWilliam et al. (1995) and Ryan et al. (1996) find -0.1 ≤ [Ni/Fe] ≤ 0.1 in the range -4 ≤ [Fe/H] ≤ 20 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars [Fe/H] Fig. 13. [X/Fe] vs. [Fe/H] for heavy elements. Symbols are the same as in Figure 11. 0. In the present work, 10 lines of Ni were used with small differences between them. Figure 12 shows that -0.13 ≤ [Ni/Fe] ≤ 0.12, in agreement with Franc¸ois et al. (2004). Copper: The analysis of Cu as a function of metallicity by Sneden & Crocker (1988) and Sneden et al. (1991) showed a linear decrease of [Cu/Fe] toward decreasing metallicities, reaching [Cu/Fe] ∼ -1 at [Fe/H] ∼ -3. This trend was confirmed by Mishenina et al. (2002), who extended the sample for halo field giants, and by Simmerer et al. (2003), whose sample in- cluded 117 giant stars in 10 globular clusters in the range of metallicities -2.4 ≤ [Fe/H] ≤ -0.8. The nucleosynthetic sites of Cu are not well established yet. Sneden et al. (1991) suggested that the Cu nucleosynthesis oc- cours mainly through the weak component of the s-process and a small contribution of the explosive burning in SN II. The ad- ditional source of Cu could be SN Ia or SN II (Matteucci et al. 1993; Timmes et al. 1995; Baraffe & Takahashi 1993). In the present work, 3 lines of Cu were used taking into account the hyperfine structure. For some stars, abundance re- sults derived from the line λ5218.2 Å are higher than those from the other 2 lines, λ5105.5 Å and λ5782.1 Å. It could be due to the gf-value, taken from a different source. However, for only one star the abundance difference between λ5218.2 Å and the other lines reaches 0.4 dex and for 3 stars, this difference is 0.25 to 0.3 dex. For the other stars, the agreement among the abundance results derived from those lines is very satisfac- tory. Except for a few stars, [Cu/Fe] is below solar for all stars D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 21 [Fe/H] Fig. 14. [X/Fe] vs. [Fe/H] for heavy elements. Symbols are the same as in Figure 11. (see Figure 12), in agreement with previous work (McWilliam 1997; Franc¸ois et al. 2004). of the s-process (Sneden et al. 1991), SN Ia (Matteucci et al. 1993) and SN II (Timmes et al. 1995). Zinc: Sneden & Crocker (1988) and Sneden et al. (1991) obtained a constant behaviour for Zn with [Zn/Fe] = 0, and this result was confirmed by Mishenina et al. (2002) for [Fe/H] > -2.0. Cayrel et al. (2004) observed an increasing trend of [Zn/Fe] toward lower metallicities. The results of Cayrel et al. (2004) are shown in Figure 5 of Franc¸ois et al. (2004), includ- ing also observations from previous work, showing that for [Fe/H] > -2.5, [Zn/Fe] is approximately constant, although with a dispersion in the range -0.25 ≤ [Zn/Fe] ≤ 0.3. Similarly to Cu, Zn can be produced by a sum of nucleosyn- thetic processes (Mishenina et al. 2002), the weak component In the present work, 4 atomic lines were used in order to compute Zn abundance. Generally, the results derived from dif- ferent lines are in good agreement. For 6 stars a larger differ- ence was observed from 0.25 to 0.6 dex, usually between the lines λ4680.1 Å and λ6362.3 Å. This difference should not be due to atomic constants, once for most stars a good agreement was found. [Zn/Fe] vs. [Fe/H] shown in Figure 12 is in good agreement with Franc¸ois et al. (2004). 22 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 1 [O/Ba] 0 -1 -2 -3 1 [Ca/Ba] 0 -1 -2 -3 1 [Cr/Ba] 0 -1 -2 -3 1 [Sc/Ba] 0 -1 -2 -3 1 [Na/Ba] 0 -1 -2 -3 0 0.5 [Mg/Ba] [Si/Ba] [Ti/Ba] [V/Ba] [Co/Ba] [Ni/Ba] [Cu/Ba] [Zn/Ba] [Al/Ba] 1 1.5 0 0.5 1 [Ba/H] 1.5 0 0.5 1 1.5 Fig. 15. [X/Ba] vs. [Ba/H]. Symbols: squares: dwarf stars (log g ≥ 3.7); triangles: subgiants (2.4 < log g < 3.7); circles: giants (log g ≤ 2.4). 4.7. s-elements In the s-elements list we included those elements that have more than 50% of s-process contribution for their abundances, according to Arlandini et al. (1999). Molibdenium was also in- cluded in this list, given that the s-process contribution for its abundance, although lower than 50%, is much larger than r- and p-processes. Details of the s-, r- and p-processes contributions for heavy elements abundances of sample stars will be found in the forthcoming paper by Allen & Barbuy (2006, paper II). on strontium are Strontium: Few studies found. Mashonkina & Gehren (2001) found -0.2 ≤ [Sr/Fe] ≤ 0.1 for disk stars in the range of metallicities -1.5 ≤ [Fe/H] ≤ 0. For one star of the sample with [Fe/H] ≈ -1 they found [Sr/Fe] ≈ 0.2. In the same range of metallicities, Gratton & Sneden (1994) found -0.2 ≤ [Sr/Fe] ≤ 0.2. Jehin et al. (1999) obtained -0.4 < [Sr/Fe] < 0 in the metallicity range -1.3 ≤ [Fe/H] ≤ -0.8. The difficulty in computing Sr abundance is the strength of the line λ4077.7 Å, which characterises barium stars. However, for most stars the abundance derived from this line was very close to that from λ4161.8 Å, as shown in Table 15. The larger D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 23 0 [O/Ba] [Mg/Ba] [Si/Ba] -1 -2 0 -1 -2 0 -1 -2 0 -1 -2 0 -1 -2 [Ca/Ba] [Ti/Ba] [V/Ba] [Cr/Ba] [Co/Ba] [Ni/Ba] [Sc/Ba] [Cu/Ba] [Zn/Ba] [Na/Ba] [Al/Ba] -1 -0.5 0 -1 -0.5 [Fe/H] 0 -1 -0.5 0 Fig. 16. [X/Ba] vs. [Fe/H]. Symbols are the same as in Figure 15. difference is for HD 147609 (0.80 dex) followed by HD 12392 (0.45 dex) and HD 188985 (0.30 dex). Lines of Sr  usually result in lower abundances than Sr  ones. This effect is also seen in the solar abundances computed by Gratton & Sneden (1994). Figure 13 shows that the relation [Sr/Fe] vs. [Fe/H] presents a larger dispersion for Sr  than for Sr . Most data are in the range 0.6 ≤ [Sr /Fe] ≤ 1.40 and 0.3 ≤ [SrI/Fe] ≤ 1.2. (1999), Tomkin & Lambert (1999) and Edvardsson et al. (1993) found an increasing trend of [Y/Fe] toward higher metallicities, ex- cluding the peculiar stars in their samples. In the present work, Y abundance was computed using synthesis of 12 lines of Yttrium: Gratton & Sneden (1994), Jehin et al. Y . In general the results derived from different lines are in good agreement. A few lines result in different abundances that can reach 0.7 dex for some stars, as can be seen in Table 15. The gf-values for the lines of Y  were those from Hannaford et al. (1982), except for the line λ6795.4 Å, with log gf from McWilliam & Rich (1994). The gf-values from Hannaford et al. (1982) were also used by Gratton & Sneden (1994) resulting in log ǫ⊙ (Y) = 2.21 ± 0.02 for the solar abun- dance, in agreement with Grevesse & Sauval (1998) value of log ǫ⊙ (Y) = 2.24 ± 0.03. Gratton & Sneden (1994) obtained - 0.3 ≤ [Y/Fe] ≤ 0.1 at [Fe/H] ≈ -1, and [Y/Fe] ≈ 0 at [Fe/H] ≈ -0.3. Tomkin & Lambert (1999) found -0.2 ≤ [Y/Fe] ≤ 0 in the 24 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars [O/Eu] [Mg/Eu] [Si/Eu] [Ca/Eu] [Ti/Eu] [V/Eu] [Cr/Eu] [Co/Eu] [Ni/Eu] [Sc/Eu] [Cu/Eu] [Zn/Eu] [Na/Eu] [Al/Eu] 1 0 -1 1 0 -1 1 0 -1 1 0 -1 1 0 -1 -0.5 0 0.5 1 -0.5 0 0.5 1 -0.5 0 0.5 1 [Eu/H] Fig. 17. [X/Eu] vs. [Eu/H]. Symbols are the same as in Figure 15. Zirconium: a combination of range -1 ≤ [Fe/H] ≤ 0. Figure 13 shows resulting values much higher for the present barium stars, in the range 0.50 ≤ [Y/Fe] ≤ 1.60. results by Burris et al. (2000, and references therein), Gratton & Sneden (1994) and Tomkin & Lambert (1999) gives a relation of [Zr/Fe] vs. [Fe/H] similar to those of Sr and Y. The dispersion increases for [Fe/H] < -1.5, whereas for higher metallicities, [Zr/Fe] is in the range -0.2 ≤ [Zr/Fe] ≤ 0.5. For the sample barium stars the disper- sion is also present, and the values are much higher than for normal stars. Table 15 shows that the abundances derived from Zr  are usually lower than those from Zr  in the range 0.40 ≤ [Zr /Fe] ≤ 1.60 and -0.20 ≤ [Zr /Fe] ≤ 1.45, the latter with higher dispersion, as shown in Figure 13. Sr, Y and Zr form the first peak of abundance of the s- process. The reason is that they have one isotope with a neutron magic number (N=50), 88Sr, 89Y and 90Zr. The larger contri- bution for the abundance of those elements comes from those isotopes. Figure 13 shows that [Sr,Y,Zr/Fe] vs. [Fe/H] are far higher than [Mo/Fe] vs. [Fe/H]. Some stars show deficiency in Mo that can reach -0.20. Molybdenum: According to Arlandini et al. (1999), the Mo abundance in the solar system has a contribution of 49.76% from s-process, 26.18% from r-process and 24.06% from p- D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 25 [O/Eu] [Mg/Eu] [Si/Eu] [Ca/Eu] [Ti/Eu] [V/Eu] [Cr/Eu] [Co/Eu] [Ni/Eu] [Sc/Eu] [Cu/Eu] [Zn/Eu] [Na/Eu] [Al/Eu] 1 0 -1 1 0 -1 1 0 -1 1 0 -1 1 0 -1 -1 -0.5 0 -1 -0.5 [Fe/H] 0 -1 -0.5 0 Fig. 18. [X/Eu] vs. [Fe/H]. Symbols are the same as in Figure 15. process. For the latter process, Mo is responsible for an abun- dance peak. of the present sample, showing a much lower pattern than the s-elements, as shown in Figure 13. In the present sample, only the line λ5570.4 Å was avail- able, for which the log gf from Bi´emont et al. (1983) was used. For the same line and log gf, Smith et al. (2000) ob- tained 1.97 for the solar abundance, close to the value by Grevesse & Sauval (1998) of 1.92 ± 0.05. Mo is little studied. Smith et al. (2000) obtained -0.21 ≤ [Mo/Fe] ≤ 0.9 for a sample of 10 red giant stars from the globular cluster ω Centauri, in the range -1.8 ≤ [Fe/H] ≤ -0.8. The range -0.20 ≤ [Mo/Fe] ≤ 1.0 was obtained for the stars Barium: Spite & Spite (1978) verified that [Ba/Fe] in- creases in the range -3 < [Fe/H] ≤ -1.5, that they called as "halo enrichment". For [Fe/H] > -1.5 the Ba enrichment is very slow, or null. This effect is confirmed by Burris et al. (2000, and references therein), Gratton & Sneden (1994), Mashonkina & Gehren (2001) and Tomkin & Lambert (1999), which show [Ba/Fe] increasing to values close to solar abun- dance in the range -4 ≤ [Fe/H] ≤ -2, and values -0.4 ≤ [Ba/Fe] ≤ 0.4 for [Fe/H] > -2. 26 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Table 18. Abundance excesses of Na, Al, α- and iron peak elements relative to Ba (upper table) and Eu (lower table). star [O/Ba] [Mg/Ba] [Si/Ba] [Ca/Ba] [Ti/Ba] [V/Ba] [Cr/Ba] [Co/Ba] [Ni/Ba] [Sc/Ba] [Cu/Ba] [Zn/Ba] [Na/Ba] [Al/Ba] [α/Ba] [Ba/H] HD 749 -0.97±0.19 -1.30±0.14 -1.01±0.12 -1.20±0.19 -1.39±0.13 -1.19±0.11 -1.28±0.11 -0.67±0.14 -1.12±0.10 -0.87±0.17 -1.13±0.14 -1.19±0.15 -1.36±0.12 -1.30±0.19 -1.01±0.29 1.12±0.20 HR 107 -1.02±0.13 -0.87±0.08 -1.01±0.08 -0.91±0.07 -0.92±0.09 -0.89±0.08 -0.98±0.07 -0.77±0.08 -1.01±0.07 -0.76±0.13 -1.12±0.11 -1.05±0.11 -0.92±0.07 -0.97±0.10 -1.03±0.13 0.59±0.07 HD 5424 -1.30±0.19 -1.28±0.14 -1.28±0.12 -1.58±0.19 -1.54±0.13 -1.65±0.11 -1.46±0.11 -1.19±0.14 -1.39±0.10 -0.99±0.17 -1.65±0.14 -1.38±0.15 -1.38±0.12 -1.33±0.19 -1.32±0.29 0.93±0.20 HD 8270 -1.03±0.13 -1.15±0.08 -1.13±0.08 -1.04±0.07 -1.24±0.09 -1.33±0.08 -1.23±0.07 -1.18±0.08 -1.17±0.07 -0.85±0.13 -1.32±0.11 -1.15±0.11 -1.06±0.07 -1.16±0.10 -1.07±0.13 0.69±0.07 HD 12392 -1.33±0.19 -1.46±0.14 -1.60±0.12 -1.52±0.19 -1.38±0.13 -1.45±0.11 -1.42±0.11 -1.04±0.14 -1.41±0.10 -1.08±0.17 -1.42±0.14 -1.49±0.15 -1.27±0.12 -1.51±0.19 -1.38±0.29 1.39±0.20 HD 13551 -0.73±0.13 -0.91±0.08 -1.06±0.08 -0.96±0.07 -1.21±0.09 -1.14±0.08 -1.22±0.07 -1.16±0.08 -1.15±0.07 -0.91±0.13 -1.33±0.11 -1.06±0.11 -1.05±0.07 -1.11±0.10 -0.78±0.13 0.72±0.07 HD 22589 -0.85±0.13 -0.67±0.08 -0.84±0.08 -0.53±0.07 -0.86±0.09 -0.96±0.08 -0.65±0.07 -0.80±0.08 -0.75±0.07 -0.72±0.13 -0.88±0.11 -0.77±0.11 -0.55±0.07 -0.78±0.10 -0.86±0.13 0.61±0.07 HD 27271 -0.69±0.19 -1.00±0.14 -0.89±0.12 -0.95±0.19 -1.12±0.13 -1.16±0.11 -1.11±0.11 -0.71±0.14 -0.85±0.10 -0.54±0.17 -0.99±0.14 -0.92±0.15 -0.66±0.12 -0.97±0.19 -0.74±0.29 0.79±0.20 HD 48565 -1.26±0.13 -1.27±0.08 -1.35±0.08 -1.24±0.07 -1.34±0.09 -1.48±0.08 -1.46±0.07 -1.37±0.08 -1.42±0.07 -1.02±0.13 -1.61±0.11 -1.26±0.11 -1.32±0.07 -1.47±0.10 -1.29±0.13 0.67±0.07 HD 76225 -1.22±0.13 -1.36±0.08 -1.38±0.08 -1.24±0.07 -1.47±0.09 -1.58±0.08 -1.47±0.07 -1.32±0.08 -1.38±0.07 -1.12±0.13 -1.47±0.11 -1.37±0.11 -1.27±0.07 -1.65±0.10 -1.26±0.13 1.04±0.07 HD 87080 -1.44±0.08 -1.57±0.08 -1.36±0.07 -1.53±0.09 -1.62±0.08 -1.58±0.07 -1.48±0.08 -1.55±0.07 -1.08±0.13 -1.73±0.11 -1.38±0.11 -1.49±0.07 -1.55±0.10 -1.49±0.13 1.04±0.07 HD 89948 -0.71±0.13 -0.92±0.08 -1.05±0.08 -0.95±0.07 -1.15±0.09 -1.09±0.08 -0.99±0.07 -1.01±0.08 -1.03±0.07 -0.91±0.13 -1.16±0.11 -0.97±0.11 -1.04±0.07 -1.29±0.10 -0.76±0.13 0.69±0.07 HD 92545 -0.71±0.13 -1.12±0.08 -1.18±0.08 -0.96±0.07 -1.16±0.09 -1.15±0.08 -1.17±0.07 -1.02±0.08 -1.13±0.07 -0.86±0.13 -1.31±0.11 -1.19±0.11 -1.00±0.07 -1.26±0.10 -0.77±0.13 0.92±0.07 HD 106191 -0.50±0.13 -0.81±0.08 -0.96±0.08 -0.78±0.07 -0.91±0.09 -0.82±0.08 -0.81±0.07 -0.73±0.08 -0.81±0.07 -0.73±0.13 -0.92±0.11 -0.80±0.11 -0.80±0.07 -0.90±0.10 -0.56±0.13 0.59±0.07 HD 107574 -1.38±0.13 -1.63±0.08 -1.70±0.08 -1.64±0.07 -1.75±0.09 -1.60±0.08 -1.86±0.07 -1.64±0.08 -1.78±0.07 -1.63±0.13 -1.93±0.11 -1.75±0.11 -1.56±0.07 -1.63±0.10 -1.43±0.13 1.16±0.07 HD 116869 -1.06±0.19 -1.06±0.14 -1.04±0.12 -1.18±0.19 -1.18±0.13 -1.29±0.11 -1.25±0.11 -0.89±0.14 -1.09±0.10 -0.89±0.17 -1.26±0.14 -1.05±0.15 -1.04±0.12 -1.02±0.19 -1.09±0.29 0.70±0.20 HD 123396 -0.88±0.19 -0.84±0.14 -1.01±0.12 -1.22±0.19 -1.37±0.13 -1.63±0.11 -1.53±0.11 -1.13±0.14 -1.24±0.10 -0.83±0.17 -1.49±0.14 -0.87±0.15 -1.18±0.12 -1.05±0.19 -0.91±0.29 0.11±0.20 HD 123585 -1.46±0.13 -1.75±0.08 -1.85±0.08 -1.72±0.07 -1.74±0.09 -1.68±0.08 -1.86±0.07 -1.45±0.08 -1.82±0.07 -1.59±0.13 -1.75±0.11 -1.69±0.11 -1.62±0.07 -1.74±0.10 -1.52±0.13 1.31±0.07 HD 147609 -1.61±0.08 -1.51±0.08 -1.60±0.07 -1.73±0.09 -1.77±0.08 -1.67±0.07 -1.58±0.08 -1.57±0.07 -0.86±0.13 -1.67±0.11 -1.43±0.11 -1.47±0.07 -1.66±0.10 -1.56±0.13 1.12±0.07 HD 150862 -0.75±0.13 -1.18±0.08 -1.16±0.08 -1.08±0.07 -1.09±0.09 -1.15±0.08 -1.10±0.07 -0.95±0.08 -1.09±0.07 -0.78±0.13 -1.15±0.11 -1.03±0.11 -1.01±0.07 -1.18±0.10 -0.81±0.13 0.93±0.07 HD 188985 -0.76±0.13 -1.21±0.08 -1.26±0.08 -1.12±0.07 -1.22±0.09 -1.15±0.08 -1.16±0.07 -1.14±0.08 -1.22±0.07 -0.96±0.13 -1.31±0.11 -1.13±0.11 -1.06±0.07 -1.32±0.10 -0.82±0.13 0.90±0.07 HD 210709 -0.75±0.19 -0.69±0.14 -0.85±0.12 -0.94±0.19 -0.98±0.13 -0.77±0.11 -0.76±0.11 -0.40±0.14 -0.79±0.10 -0.55±0.17 -0.90±0.14 -0.89±0.15 -0.79±0.12 -0.75±0.19 -0.78±0.29 0.69±0.20 HD 210910 -0.55±0.19 -0.88±0.14 -0.82±0.12 -1.44±0.19 -1.50±0.13 -1.27±0.11 -1.67±0.11 -0.85±0.14 -1.01±0.10 -0.71±0.17 -1.15±0.14 -0.81±0.15 -1.16±0.12 -0.82±0.19 -0.60±0.29 0.62±0.20 -0.95±0.13 -1.25±0.08 -1.41±0.08 -1.29±0.07 -1.43±0.09 -1.24±0.08 -1.52±0.07 -1.41±0.08 -1.40±0.07 -1.16±0.13 -1.57±0.11 -1.34±0.11 -1.22±0.07 -1.43±0.10 -1.01±0.13 0.75±0.07 HD 222349 BD+18 5215 -1.03±0.13 -1.29±0.08 -1.41±0.08 -1.39±0.07 -1.50±0.09 -1.37±0.08 -1.46±0.07 -1.11±0.08 -1.44±0.07 -1.28±0.13 -1.55±0.11 -1.41±0.11 -1.29±0.07 -1.56±0.10 -1.08±0.13 0.93±0.07 HD 223938 -0.87±0.19 -1.13±0.14 -1.19±0.12 -1.16±0.19 -1.38±0.13 -1.41±0.11 -1.40±0.11 -1.09±0.14 -1.21±0.10 -0.90±0.17 -1.37±0.14 -1.15±0.15 -1.18±0.12 -1.30±0.19 -0.92±0.29 0.87±0.20 [Eu/H] [Mg/Eu] [Cu/Eu] [Na/Eu] [Cr/Eu] [Co/Eu] [Si/Eu] [Ca/Eu] [Ni/Eu] [Sc/Eu] [Ti/Eu] [V/Eu] [Al/Eu] [α/Eu] ... ... star [O/Eu] [Zn/Eu] ... HD 749 HR 107 HD 5424 HD 8270 HD 12392 HD 13551 HD 22589 HD 27271 HD 48565 HD 76225 HD 87080 HD 89948 HD 92545 HD 106191 HD 107574 HD 116869 HD 123396 HD 123585 HD 147609 HD 150862 HD 188985 HD 210709 HD 210910 HD 222349 BD+18 5215 0.19±0.14 -0.07±0.11 -0.19±0.10 -0.17±0.09 -0.28±0.11 -0.15±0.10 -0.24±0.10 HD 223938 0.18±0.16 -0.27±0.12 -0.02±0.18 -0.28±0.15 -0.34±0.16 -0.51±0.13 -0.45±0.20 -0.16±0.30 0.27±0.21 -0.12±0.20 -0.45±0.15 -0.16±0.13 -0.35±0.20 -0.54±0.14 -0.34±0.13 -0.43±0.12 -0.11±0.14 0.04±0.11 -0.10±0.10 0.14±0.10 -0.10±0.10 0.15±0.14 -0.21±0.13 -0.14±0.13 -0.01±0.10 -0.06±0.12 -0.12±0.14 -0.32±0.10 0.00±0.09 -0.01±0.11 0.02±0.10 -0.07±0.10 0.03±0.18 -0.63±0.15 -0.36±0.16 -0.36±0.13 -0.31±0.20 -0.30±0.30 -0.09±0.21 -0.28±0.20 -0.26±0.15 -0.26±0.13 -0.56±0.20 -0.52±0.14 -0.63±0.13 -0.44±0.12 -0.17±0.16 -0.37±0.12 -0.24±0.14 -0.36±0.11 -0.34±0.10 -0.25±0.09 -0.45±0.11 -0.54±0.10 -0.44±0.10 -0.39±0.10 -0.38±0.10 -0.06±0.14 -0.53±0.13 -0.36±0.13 -0.27±0.10 -0.37±0.12 -0.28±0.14 -0.10±0.10 -0.30±0.20 -0.43±0.15 -0.57±0.13 -0.49±0.20 -0.35±0.14 -0.42±0.13 -0.39±0.12 -0.01±0.16 -0.38±0.12 -0.05±0.18 -0.39±0.15 -0.46±0.16 -0.24±0.13 -0.48±0.20 -0.35±0.30 0.36±0.21 0.04±0.14 -0.38±0.13 -0.11±0.13 -0.10±0.10 -0.16±0.12 0.22±0.14 0.17±0.14 -0.23±0.10 0.04±0.11 -0.11±0.10 -0.01±0.09 -0.26±0.11 -0.19±0.10 -0.27±0.10 -0.21±0.10 -0.20±0.10 -0.18±0.14 0.00±0.11 -0.17±0.10 0.14±0.09 -0.19±0.11 -0.29±0.10 0.02±0.10 -0.13±0.10 -0.08±0.10 -0.05±0.14 -0.21±0.13 -0.10±0.13 0.12±0.10 -0.11±0.12 -0.19±0.14 -0.06±0.10 -0.12±0.20 -0.43±0.15 -0.32±0.13 -0.38±0.20 -0.55±0.14 -0.59±0.13 -0.54±0.12 -0.14±0.16 -0.28±0.12 0.03±0.18 -0.42±0.15 -0.35±0.16 -0.09±0.13 -0.40±0.20 -0.17±0.30 0.22±0.21 -0.32±0.14 -0.33±0.11 -0.41±0.10 -0.30±0.09 -0.40±0.11 -0.54±0.10 -0.52±0.10 -0.43±0.10 -0.48±0.10 -0.08±0.14 -0.67±0.13 -0.32±0.13 -0.38±0.10 -0.53±0.12 -0.35±0.14 -0.27±0.10 -0.12±0.14 -0.26±0.11 -0.28±0.10 -0.14±0.09 -0.37±0.11 -0.48±0.10 -0.37±0.10 -0.22±0.10 -0.28±0.10 -0.02±0.14 -0.37±0.13 -0.27±0.13 -0.17±0.10 -0.55±0.12 -0.16±0.14 -0.06±0.10 -0.62±0.11 -0.75±0.10 -0.54±0.09 -0.71±0.11 -0.80±0.10 -0.76±0.10 -0.66±0.10 -0.73±0.10 -0.26±0.14 -0.91±0.13 -0.56±0.13 -0.67±0.10 -0.73±0.12 -0.67±0.14 0.22±0.10 0.12±0.14 -0.09±0.11 -0.22±0.10 -0.12±0.09 -0.32±0.11 -0.26±0.10 -0.16±0.10 -0.18±0.10 -0.20±0.10 -0.08±0.14 -0.33±0.13 -0.14±0.13 -0.21±0.10 -0.46±0.12 0.07±0.14 -0.14±0.10 0.01±0.14 -0.40±0.11 -0.46±0.10 -0.24±0.09 -0.44±0.11 -0.43±0.10 -0.45±0.10 -0.30±0.10 -0.41±0.10 -0.14±0.14 -0.59±0.13 -0.47±0.13 -0.28±0.10 -0.54±0.12 -0.05±0.14 0.20±0.10 0.18±0.14 -0.13±0.11 -0.28±0.10 -0.10±0.09 -0.23±0.11 -0.14±0.10 -0.13±0.10 -0.05±0.10 -0.13±0.10 -0.05±0.14 -0.24±0.13 -0.12±0.13 -0.12±0.10 -0.22±0.12 0.12±0.14 -0.09±0.10 -0.14±0.14 -0.39±0.11 -0.46±0.10 -0.40±0.09 -0.51±0.11 -0.36±0.10 -0.62±0.10 -0.40±0.10 -0.54±0.10 -0.39±0.14 -0.69±0.13 -0.51±0.13 -0.32±0.10 -0.39±0.12 -0.19±0.14 -0.08±0.10 -0.20±0.20 -0.20±0.15 -0.18±0.13 -0.32±0.20 -0.32±0.14 -0.43±0.13 -0.39±0.12 -0.03±0.16 -0.23±0.12 -0.03±0.18 -0.40±0.15 -0.19±0.16 -0.18±0.13 -0.16±0.20 -0.23±0.30 -0.16±0.21 -0.08±0.20 -0.04±0.15 -0.21±0.13 -0.42±0.20 -0.57±0.14 -0.83±0.13 -0.73±0.12 -0.33±0.16 -0.44±0.12 -0.03±0.18 -0.69±0.15 -0.07±0.16 -0.38±0.13 -0.25±0.20 -0.11±0.30 -0.69±0.21 -0.50±0.14 -0.79±0.11 -0.89±0.10 -0.76±0.09 -0.78±0.11 -0.72±0.10 -0.90±0.10 -0.49±0.10 -0.86±0.10 -0.63±0.14 -0.79±0.13 -0.73±0.13 -0.66±0.10 -0.78±0.12 -0.56±0.14 0.35±0.10 -0.78±0.11 -0.68±0.10 -0.77±0.09 -0.90±0.11 -0.94±0.10 -0.84±0.10 -0.75±0.10 -0.74±0.10 -0.03±0.14 -0.84±0.13 -0.60±0.13 -0.64±0.10 -0.83±0.12 -0.73±0.14 0.29±0.10 0.05±0.14 -0.32±0.13 -0.20±0.13 -0.18±0.10 -0.35±0.12 0.08±0.14 -0.35±0.11 -0.33±0.10 -0.25±0.09 -0.26±0.11 -0.32±0.10 -0.27±0.10 -0.12±0.10 -0.26±0.10 0.02±0.14 0.10±0.10 0.15±0.14 -0.30±0.11 -0.35±0.10 -0.21±0.09 -0.31±0.11 -0.24±0.10 -0.25±0.10 -0.23±0.10 -0.31±0.10 -0.05±0.14 -0.40±0.13 -0.22±0.13 -0.15±0.10 -0.41±0.12 0.09±0.14 -0.01±0.10 0.10±0.18 -0.25±0.15 -0.24±0.16 -0.14±0.13 -0.10±0.20 -0.13±0.30 0.04±0.21 -0.10±0.20 -0.04±0.15 -0.20±0.13 -0.29±0.20 -0.33±0.14 -0.12±0.13 -0.11±0.12 0.25±0.16 -0.14±0.12 -0.10±0.20 -0.43±0.15 -0.37±0.13 -0.99±0.20 -1.05±0.14 -0.82±0.13 -1.22±0.12 -0.40±0.16 -0.56±0.12 -0.26±0.18 -0.70±0.15 -0.36±0.16 -0.71±0.13 -0.37±0.20 -0.15±0.30 0.17±0.21 0.19±0.14 -0.11±0.11 -0.27±0.10 -0.15±0.09 -0.29±0.11 -0.10±0.10 -0.38±0.10 -0.27±0.10 -0.26±0.10 -0.02±0.14 -0.43±0.13 -0.20±0.13 -0.08±0.10 -0.29±0.12 0.13±0.14 -0.39±0.10 0.14±0.14 -0.29±0.10 0.11±0.10 -0.22±0.10 -0.06±0.14 -0.33±0.13 -0.19±0.13 -0.07±0.10 -0.34±0.12 -0.02±0.20 -0.28±0.15 -0.34±0.13 -0.31±0.20 -0.53±0.14 -0.56±0.13 -0.55±0.12 -0.24±0.16 -0.36±0.12 -0.05±0.18 -0.52±0.15 -0.30±0.16 -0.33±0.13 -0.45±0.20 -0.07±0.30 0.02±0.21 ... In the present work, all values of [Ba/Fe] are in the range of 0.8 ≤ [Ba/Fe] ≤ 1.80, showing high Ba overabundance relative to Fe, which is a defining characteristics of barium stars. Five lines of Ba were used taking into account hyperfine structure. In general, good agreement on abundances derived from different lines was found, as shown in Table 15. Figures 15 and 16 and Table 18 show abundances of α- and iron peak elements rela- tive to Ba. The abundance excesses relative to Ba show a de- creasing trend toward increasing [Ba/H]. The decreasing trend of [iron peak/Ba] vs. [Ba/H] could be explained by the sec- ondary charater of the s-process, where iron peak elements are the seed nuclei. [X/Ba] relative to [Fe/H] show an increasing trend with increasing [Fe/H]. This trend is in agreement with the larger efficiency of the s-process in producing heavy ele- ments such as Ba, at lower metallicities. The s-process site is different from that of Al, Na, α- and iron peak elements, there- fore it is not surprising that [X/Ba] vs. [Ba/H] and [X/Ba] vs. [Fe/H] are not constant. In both figures 15 and 16, the halo gi- ant star HD 123396 ([Fe/H] = -1.19) is out of the trend shown by the other stars. Lanthanum: Burris et al. (2000, and references therein) and Gratton & Sneden (1994) show a similar behaviour for [La/Fe] and [Ba/Fe]. Gratton & Sneden (1994) found a range of -0.4 ≤ [La/Fe] ≤ 0.05 at -2 ≤ [Fe/H] ≤ 0, and Burris et al. (2000), in the range -2 ≤ [Fe/H] ≤ -0.5, found 0 ≤ [La/Fe] ≤ 0.5, with 2 stars with [La/Fe] ≤ 0. In the present work, 8 lines of La  were used taking into account the hyperfine structure. For some stars abundance dif- ferences between lines can reach 0.7 dex, mainly involving the line λ4086.7 Å. The origin of this difference is not related to the atomic constants since for several stars the agreement is very good. As an example, the abundance obtained from this line for HD 5424 was lower than from the line λ5797.6 Å, while for HD 22589 the result is inverted. If the problem were in the atomic constants, the difference pattern would be always the same. The range obtained was 0.6 ≤ [La/Fe] ≤ 1.70 (see Figure D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 27 13), showing the large La overabundance relative to Fe for bar- ium stars, in contrast to normal stars. Neodymium: Gratton & Sneden Cerium: Gratton & Sneden (1994) and Jehin et al. (1999) showed for [Ce/Fe] a behaviour similar to [Ba/Fe] and [La/Fe]. Their results are in the range -0.4 ≤ [Ce/Fe] ≤ 0.15 for -2 ≤ [Fe/H] ≤ 0. In the present work, 8 lines of Ce  were used for abundance determination. As for La, in some cases, the abun- dance result of one line was very different from the others, but the effect on the average is low, as can be seen in Table 15, 16 and 17. Abundance excess relative to Fe is in the range 0.4 ≤ [Ce/Fe] ≤ 1.80, with large dispersion among the stars. (1994), Tomkin & Lambert (1999), Jehin et al. (1999) and Burris et al. (2000, and references therein) show that [Nd/Fe] behaves similarly to Ba, La and Ce. According to Figure 5 from Burris et al. (2000), a high dispersion in the range -3 ≤ [Fe/H] ≤ -1.5 is present, with low values close to [Nd/Fe] ≈ -0.6 and high values of ≈ +1.5. For [Fe/H] > -1.5 the results are in the range -0.3 ≤ [Nd/Fe] ≤ +0.3. In the calculation of Nd abundance for the present sam- ple, the hyperfine structure was neglected. According to Hartog et al. (2003), only the odd isotopes 143Nd and 145Nd show hyperfine structure, accounting for 20.5% of the abun- dance, hence the effect can be ignored. The abundances derived from 9 lines of Nd  were computed, and they show a good agreement, with a few exceptions, similar to previous elements described here. The range obtained for the abundance excess relative to Fe was 0.3 ≤ [Nd/Fe] ≤ 1.70. Ba, La, Ce and Nd form the second abundance peak of the s-process because of nuclides 138Ba, 139La, 140Ce and 142Nd are neutron magic nuclei (N=82). Lead: There are a few stars for which the lead abundance was determined. Van Eck et al. (2003), show that some CH stars show high Pb abundances, and Sivarani et al. (2004) gath- ered about 30 halo stars with high Pb abundances. There are 4 Pb isotopes: 204Pb, 206Pb, 207Pb and 208Pb. The last is dou- bly magical (neutron magic in N=126 and proton magic in Z=82) being the responsible for the third abundance peak of the s-process. The best lines for the abundance calculation are λ3639.6 Å, λ3683.5 Å, λ3739.9 Å, λ4057.8 Å and λ7229 Å. Wavelengths λ < 4000 Å are not reliable with the FEROS spec- tra and the λ7229 Å line is too weak. For these reasons, the Pb abundance was determined using the line λ4057.8 Å, for which blends have been taken into account. The results are in the range -0.2 ≤ [Pb/Fe] ≤ 1.6, showing a large dispersion. According to Arlandini et al. (1999), the s-process is respon- sible for 46% of the Pb abundance with no contribution from the r-process. The missing abundance is generally attributed to the strong component of the s-process. 4.8. r-elements We included in the r-elements list only those with more than 50% of r-process contribution for their abundances, according to Arlandini et al. (1999). Europium: (2000), Jehin et al. Gratton & Sneden (1994), (1999), Woolf et al. Burris et al. and (1995) Mashonkina & Gehren (2001) provide an increasing lin- ear relation between log ǫ(Eu) and [Fe/H]. On the other hand, [Eu/Fe] decreases toward higher metallicities. In the present work the lines λ4129.7 Å, λ4205 Å, λ6437.7 Å and λ6645.1 Å were used for computing Eu abundances, tak- ing into account the hyperfine structure. The abundances de- rived from these 2 first lines are usually lower than those de- rived from the 2 last lines (see Table 15). The 2 first lines show a blend with CN lines in the cooler stars, whereas for the hotter stars, this blend is negligible. Figure 17 and Table 18 show the relation between [Eu/H] and abundance excess of α- and iron peak elements relative to Eu, showing a larger dispersion than Figure 15. With the excep- tion of O, Mg, Co and Sc, the other α- and iron peak elements show [X/Eu] ≤ 0. It seems that there is an interval of [Eu/H] where [X/Eu] is constant and, for [Eu/H] > 0 a decreasing trend is seen, that could reflect the secondary character of the r-process with the iron peak elements as the seed nuclei. A pri- mary r-process (e.g. Meyer 1994) contribution would form Eu without depletion of other elements. Relative to [Fe/H] (Figure 18), [X/Eu] seems to be constant, with the possible exception of [Cr/Eu], [Co/Eu] and [Cu/Eu] which seem to increase to- ward increasing metallicities. The constancy of [X/Fe] could be due to the fact that α-elements, and probably Na and Al, are produced in massive stars (mostly in hydrostatic phases), and they are released in their SNae type II explosions as well as the r-elements. Thus, if all of them were released through the same events, their ratios are expected to be constant. The dispersion comes from the fact that stars of different masses produce different amounts of each element. Regarding the in- creasing or constant behaviour of [iron peak/Eu] vs. [Fe/H], it can be due to their production in different amounts in SNae II and SNae Ia. A larger range of metallicities expanded toward lower values could illustrate better such behaviour. Praseodymium: Little is done on Pr. Gratton & Sneden (1994) determined Pr abundance for metal-poor stars and ob- tained results in the range -0.2 ≤ [Pr/Fe] ≤ 0.3, for -1.5 ≤ [Fe/H] ≤ 0. Figure 14 shows that the present results are in the range 0.2 ≤ [Pr/Fe] ≤ 1.40, showing the contribution of the s-process to the Pr abundance. The three lines used were in good internal agreement, with a few exceptions, as shown in Table 15. Samarium: Gratton & Sneden (1994) included Sm in their analysis with results in the range -0.3 ≤ [Sm/Fe] ≤ 0.2 for -1.5 ≤ [Fe/H] ≤ 0. In the present work, the results are in the range 0 ≤ [Sm/Fe] ≤ 1.40 (Figure 14). The Sm lines are very weak and the hfs of the 5 lines can be neglected. The abundances derived from several lines are in good agreement, as can be seen in Table 15. Gadolinium: Gadolinium abundances are essentially not found in the literature, at least, in the range of metallicities of the present sample. The lines are very weak or even invisible in the spectra. Furthermore, they are only present at λ < 5000 Å with several blends. In the present work 3 lines of Gd were used. For most of them there was good agreement among the abundances derived from different lines (see Table 15). Figure 14 shows low dis- persion in the relation [Gd/Fe] vs. [Fe/H]. 28 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Dysprosium: Gratton & Sneden (1994) studied stars in the range of metallicities -1.5 ≤ [Fe/H] ≤ 0, and they included Dy in their analysis. They found a range of -0.2 ≤ [Dy/Fe] ≤ 0.2 except for one star for which they found [Dy/Fe] ≈ -0.4. The results of the present sample are in the range of -0.25 ≤ [Dy/Fe] ≤ 1.65, shown in Figure 14. Two lines were available in the spectra of the sample stars, from which a good agreement was obtained. 4.9. Uncertainties The abundance uncertainties were calculated by verifying how much the variation of each input parameter changes the out- put value log Ap. Table 19 shows the values taken into account in this calculation and the resulting uncertainties for each ele- ment. The procedure was adopted for 2 stars, HD 5424 of low log g = 2.0 and HD 150862 of high log g = 4.6. The uncertainty on the output value is given by σAp = q(∆AT )2 + (∆Alg)2 + (∆Av)2 + (∆Am)2 (14) where ∆AT , ∆Alg, ∆Av and ∆Am are the differences in Ap due to variations of 1σ in the temperature, log g, microturbulent velocity and metallicity, respectively. The average value of Ap (Apm) is obtained by averaging individual abundances of several lines, and not from several measurements of the same line. In the latter case, the standard deviation could be used to calculate the uncertainty on Apm. Considering this, we found more suitable to apply propagation of errors taking into account the uncertainty calculated with equation 14. Thus, the uncertainty on Apm is σApm = σAp √n (15) where n is the number of lines. The uncertainty on the loga- rithm of Apm is σlog(Apm) = σApm Apm ln 10 . (16) The abundance log ǫ(X) is related to the output of the syn- thesis program by log ǫ(X) = log Apm + [Fe/H]. Therefore, the uncertainty is σlog ǫ(X) = qσ2 log (Apm) + σ2 [Fe/H]. (17) The relation between the abundance excess relative to iron [X/Fe] and the output value of the synthesis program is [X/Fe] = log Apm - log ǫ(X) is the solar abundance of the element "X". The uncertainty is calculated by , where log ǫ(X) ⊙ ⊙ σ[X/Fe] = qσ2 log(Apm) + σ2 logǫ⊙ (X) (18) The uncertainty on [α/Fe], which contains the contribution of the uncertainty on the abundance of each element taken into account in the calculation of the α's, is given by σ[α/Fe] = qσ2 log ǫ(α) + σ2 (α) + σ2 [Fe/H] (19) log ǫ⊙ with σlog ǫ(α) = 1 nǫ(α) vt n Xi=1 (10log ǫ(Xi))2σ2 log ǫ(Xi) (20) where n is the number of elements "X" used in the calculation of α, ǫ(α) = 10log ǫ(α). In this work, n=5 for most stars. For 2 stars, HD 87080 and HD 147609, n=4 because the oxygen was excluded from the calculation. The σlog ǫ(α) is similar to ⊙ expression 20. For the iron peak elements, the uncertainty is similar to ex- pression 19. In Figures 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 17, 18 are shown the maximum value of uncertainties on each axis. In the corre- sponding tables it is possible to check all the values. 5. Conclusions The barium stars have been studied throughout more than five decades, however several open questions still remain relative to their origin and characteristics. The aim of the present work was to increase the knowledge about this class of peculiar stars. For the first time a detailed study on the behaviour of abun- dance ratios for a large number of elements is presented for a relatively large sample of barium stars. As the first outcome of this work, the results of the atmo- spheric parameters show that the sample consisted of stars of different luminosity classes with 4300 ≤ Te f f ≤ 6500 K and 1.4 ≤ log g ≤ 4.6. The metallicities obtained are typical of barium stars, in the range -1.2 ≤ [Fe/H] ≤ 0.0, most of them with -0.62≤ [Fe/H] ≤ 0.0. For 7 stars a significant difference was found between the metallicities resulting from Fe  and Fe  lines, (∆[Fe/H] ≥ 0.2 dex). It is very important this to be understood in order to be possible to determine reliable abundances, given that the er- rors in metallicities affect the resulting abundances. The differ- ence appears when one determines the surface gravities through the classical equation, which requires the mass values. In the present work, two methods were adopted. In the first case, the masses were derived from isochrones and then the surface grav- ities were determined. In the second case, the surface gravities were derived from ionization equilibrium and then the masses. The difference between the surface gravities determined from the two methods can reach 0.6 dex and in the worst case, over 1 dex for the star HD 147609. This difference reflects in the differences between the metallicities derived from lines of Fe  and Fe . One point to discuss is the fact that the masses de- termined using the log g from ionization equilibrium are very small for these 7 stars. Another point is that ∆[Fe/H] depends on the gf-values adopted (see Sect. 3.5). Fe  lines are less af- fected by NLTE effects than Fe , and more accurate gf-values for Fe  lines were used (Mel´endez & Barbuy 2006), reducing the difference in log g by 0.2 dex. There is a good agreement between the present results and literature data. The abundance results obtained for the sample stars show that there are no correlations with the luminosity classes. The abundances found for the α-, iron peak, Li, Al and D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 29 Na are compatible with the values of [X/Fe] given in the liter- ature for normal disk stars in the same range of metallicities, and the s-overabundance is independent of luminosity class. There are not enough halo stars in the present sample to iden- tify differences between halo and disk barium stars. The range of metallicities is too small to allow a well-defined trend in the [X/Fe] vs. [Fe/H] of α- and iron peak elements. For heavy el- ements, there is a small variation that can be explained by the variable amount of enriched material that each star received from the more evolved companion. The Li abundance decreases toward lower temperatures. This result is consistent with the discussion in the literature that the Li is depleted along the red giant branch evolution. Less evolved stars show higher C abundances, and [C/O] is approximately constant with metallicity. Besides being C-rich, the barium stars of the present sample are N-rich. It happens for all stars including the less evolved ones, suggesting that N is also responsible for the CNO excess in these stars. For most stars, the excesses of Na, Al, Mg, Si and Ca rel- ative to Fe are within -0.2 < [X/Fe] < 0.2. O reaches higher values and Ti, lower values. [Ti/Fe] are approximately similar to those of [V/Fe], [Cr/Fe] and [Ni/Fe], identifying Ti more as iron peak than an α-element. For some stars the odd-even ef- fect, where Mg and Si are overabundant relative to Na and Al, can be observed, however, for several stars the abundances of Na and Al are higher than those of Si and Mg. Among iron peak elements, the Sc has the highest abun- dances. This result is in agreement with Nissen et al. (2000), that identified an "α-element" behaviour for Sc. For Co, the theoretical prediction is [Co/Fe] < 0 for the range of metallici- ties of the present sample, however, for most stars [Co/Fe] > 0 was obtained. For other iron peak elements, V, Cr, Ni and Zn, [X/Fe] show a lower range of values than [Co/Fe]. Except for 4 stars, [Cu/Fe] is below solar for the present sample. The excesses of Na, Al, α- and iron peak elements relative to Ba show a decreasing trend with [Ba/H], whereas [X/Ba] vs. [Fe/H] show an increasing trend. Considering that Ba repre- sents the s-process elements, one can consider that these corre- lations describe the relations between s-process and other nu- cleosynthetic processes. Regarding the r-process element Eu, there is a range of [Eu/H] where [X/Eu] is essentially constant, [X/Eu] showing a decreasing trend toward higher [Eu/H]. For most stars, [X/Eu] ≤ 0, except for O, Mg, Co and Sc. [X/Eu] vs. [Fe/H] is constant for Na, Al and α-elements as expected. [X/Ba] and [X/Eu] for the sample stars characterises the abundance behaviour of dif- ferent elements relative to the s- and r-processes. Acknowledgements. We acknowledge partial financial support from the Brazilian Agencies CNPq and FAPESP. DMA acknowledges a FAPESP PhD fellowship n◦ 00/10405-8 and a FAPERJ post-doctoral fellowship n◦ 152.680/2004. We are grateful to Anita G´omez for suggesting the analysis of Hipparcos dwarf barium stars candidates, to Andrew McWilliam for making available his code on hyper- fine structure and to the referee, Nils Ryde, for useful comments. We are also grateful to Licio da Silva, Luciana Pomp´eia, Paula Coelho and Jorge Mel´endez for carrying out some observations of our sample spectra, and to Gustavo Porto de Mello, Wladimir and Graziela, for helping us with photometric observations. This pub- lication makes use of the SIMBAD database and of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. References Allen D.M., Barbuy.B., 2006, A&A, submitted (paper II) Allende Prieto C., Garcia-Lopez R., Lambert D.L., Gustafsson B. 1999, ApJ, 527, 879 Allende Prieto C., Lambert D.L., Asplund M., 2001, ApJ 556, L63 Alonso A., Arribas S., Martinez-Roger C., 1995, A&A, 297, 197 Alonso A., Arribas S., Martinez-Roger C. 1996, A&A 313, 873 Alonso A., Arribas S., Martinez-Roger C. 1999, A&AS 140, 261 Andersson H., Edvardsson B. 1994, A&A, 290, 590 Arlandini C., Kappeler F., Wisshak K. 1999, ApJ, 525, 886 Arnett W.D. 1971, ApJ, 166, 153 Asplund M., Grevesse N., Sauval J. 2005, astro-ph/0410214 Baraffe I., Takahashi K. 1993, A&A, 280, 476 Barbuy B., 1988, A&A, 191, 121 Barbuy B., Jorissen A., Rossi S.C.F., Arnould M. 1992, A&A, 262, 216 Barbuy B., Perrin, M.-N., Katz, D., Coelho, P., Cayrel, R., Spite, M., Van't Veer-Menneret, C., 2003, A&A, 404, 661 Barbuy B., Renzini A., Ortolani S., Bica E., Guarnieri M.D., 1999, A&A, 341, 539 Barklem P.S., O'Mara B.J., Ross J.E., 1998, MNRAS, 296, 1057 Barklem P.S., Piskunov, N., O'Mara, B. J., 2000, A&AS, 142, 467 Baumuller D., Butler K., Gehren T. 1998, A&A, 338, 637 Baumuller D., Gehren T. 1997, A&A, 325, 1088 Bell R.A., Eriksson K., Gustafsson B., Nordlund A., 1976, A&AS 23, 37 Bergstrom H., Bi´emont E., Lundberg H., Persson A., 1988, A&A, 192, 337 Bertelli G., Bressan A., Chiosi C., Fagotto F., Nasi E., 1994, A&AS, 106, 275 Bessell M.S. 1979, PASP 91, 589 Bessell M.S., Brett J.M. 1988, PASP 100, 1134 Bessell M.S., Castelli, F., Plez, B. 1998, A&A 333, 231 Bidelman W.P., Keenan P.C. 1951, ApJ 114. 473 Biehl D., 1976, Ph.D. Thesis, Kiel University Bielski, A. 1975, JQSRT, 15, 463 Bi´emont E. Gamir H.P., Palmeri P., Li Z.S., Svanberg S., 2000, MNRAS, 312, 116 Bi´emont E., Godefroid M., 1980, A&A, 84, 361 Bi´emont E., Grevesse N., Hannaford P., Lowe R.M., 1981, ApJ, 248, 867 Bi´emont E, Grevesse N., Hannaford P., Lowe R.M. 1989, A&A, 222, 307 Bi´emont E., Grevesse N., Hannaford P., Lowe R.M., Whaling W. 1983, ApJ, 275, 889 30 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Bi´emont E., Lowe R.M. 1993, A&A, 273, 665 Boesgaard A.M. 1970, ApJ, 161, 1003 Bond H.E., Neff I.S., 1969, ApJ, 158, 1235 Bond H.E., Pollacco D.L., Webbink R. 2003, ApJ, 125, 260 Bohm-Vitense E. 1980, ApJ, 239, L79 Bohm-Vitense E., Carpenter K., Robinson R., Ake T., Brown J. 2000, ApJ, 533, 969 Bohm-Vitense E., Johnson H.R. 1985 ApJ, 293, 288 Brown J.A., Sneden C., Lambert D.L., Dutchover E. 1989, ApJS, 71, 293 Burris, D.L., Pilachowski C.A., Armandroff T.E., Sneden C., Hakkila J., 1989, A&A, 213, 204 Hannaford P., Lowe R.M., Grevesse N., Bi´emont E., Whaling W. 1982, ApJ, 261, 736 Hannaford P., Lowe R.M., 1983, Opt. Eng, 22, 532 Hartog E.A.D., Lawler J.E., Sneden C., Cowan J.J. 2003, ApJS, 148, 543 Jacoby G.H. 1981, ApJ, 244, 903 Jasniewicz G., Th´evenin F., Monier R., Skiff B. 1996, A&A, 307, 200 Jeffries R.D., Smalley B. 1996, A&A, 315, L19 Jehin E., Magain P., Neuforge C., Noels A., Parmentier G., Cowan J.J., Roe H. 2000, ApJ, 544, 302 Thoul A.A. 1999, A&A, 341, 241 Cardelli J.A., Clayton G.C., Mathis J.S. 1989, ApJ 345, 245 Carpenter J.M. 2001, AJ 121, 2851 Castilho B.V., Gregorio-Hetem J., Spite F., Spite M., Barbuy B. 1998, A&AS, 127, 139 Castilho B.V., Pasquini L., Allen D.M., Barbuy B., Molaro P., 2000, A&A, 361, 92 Catchpole R.M., Robertson B.S.C., Warren P.R., 1977, MNRAS, 181,391 Cayrel, R., 1989, em The impact of very high S/N spectroscopy on stellar physics, ed. G. Cayrel de Strobel, & M. Spite (Dordrecht: Kluwer Academic Publ.), IAU Symp., 132, 345 Cayrel, R., Depagne E., Spite M., et al. 2004, A&A, 416, 1117 Cayrel, R., Jugaku, 1963, AnAp, 26, 495 Cayrel, R., Perrin M.N., Barbuy B., Buser R., 1991, A&A, 247, 108 Chen, B., Vergely J.L., Valette B., Carraro G. 1998, A&A 336, 137 Clegg R.E.S., Lambert D.L., Tomkin J. 1981, ApJ, 250, 262 Coelho P., Barbuy, B., Mel´endez, J., Schiavon, R.P., Castilho, B.V. 2005, A&A, 443, 735 Corliss C.H., Bozman W.R., 1962, Cram L. 1999 em Transactions of International Astronomical Union, Volume XXIIIB, 141, Editor: J. Andersen the Kipper T., Wallerstein G. 1990, PASP, 102, 574 Kraft R.P., Ivans I.I. 2003, PASP, 115, 143 Kurucz R. L., Furenlid I., Brault J., 1984, Solar flux atlas from 296 to 1300 nm, National Solar Observatory Atlas, Sunspot (New Mexico: National Solar Observatory) Kusz J., 1992, A&AS, 92, 517 Lage C.S., Whaling W. 1976, JQSRT, 16, 537 Lambert D.L., 1978, MNRAS, 182, 249 Lawler J.E., Bonvallet G., Sneden C. 2001a, ApJ, 556, 452 Lawler J.E., Dakin J.T. 1989, J.Opt.Soc.Am., B6, 1457 Lawler J.E., Wickliffe M.E., Hartog A.D. 2001b, ApJ, 563, 1075 Lejeune T., Cuisinier F., Buser R. 1998, A&AS 130, 65 Liang Y.C., Zhao G., Chen Y.Q., Qiu H.M., Zhang B. 2003, A&A, 397, 257 Lodders K., 2003, ApJ, 591, 1220 Lu, P.K. 1991, AJ, 101, 2229 Luck R.E., Bond H.E., 1991, ApJS, 77, 515 Maier R.S., Whaling W., 1977, JQSRT, 18, 501 Martin, W.C., Fuhr, J.R., Kelleher, D.E., et al. 2002, NIST Atomic Database (version 2.0), http://physics.nist.gov/asd. National Technology, Institute Gaithersburg, MD. Standards of and Martin G.A., Fuhr J.R., Weise W.L. 1988, J.Phys.Chem.Ref.D, Cutri R.M., Skrutskie M.F., van Dyk S., et al. 2003, VizieR Vol 17, No. 3 Online Data Catalog, 2246 Denn G.R., Luck R.E., Lambert D.L. 1991, ApJ, 377, 657 Dominy J.F., Lambert D.L. 1983, ApJ, 270, 180 Domogatskii G.V., Eramzhyan R.A., Nadyozhin D.K. 1977, em Neutrino 77 (Nauka, Moscow), p. 115 Edvardsson B., Andersen J., Gustafsson B., Lambert D.L., Nissen P.E., Tomkin J. 1993, A&A, 275, 101 Franc¸ois P., Matteucci F., Cayrel R., Spite M., Spite F., Chiappini C. 2004, A&A, 421, 613 Fuhrmann K., Axer M., Gehren T., 1995, A&A, 301, 492 Goly A., Kusz J., Nguyen Quang B., Weniger S., 1991, JQSRT, 45, 157 G´omez A.E., Luri X., Grenier S., Pr´evot L., Mennessier M.O., Figueras F., Torra J. 1997, A&A 319, 881 Goswami, A., Prantzos, N. 2000, A&A, 359, 191 Gratton R.G., Carreta E., Castelli F., 1996, A&A, 314, 191 Gratton R.G., Sneden C. 1991, A&A, 241, 501 Gratton R.G., Sneden C. 1994, A&A, 287, 927 Grevesse N., Sauval A.J. 1998, Space Sci. Rev., 85, 161 Gustafsson B., Bell K.A., Eriksson K., Nordlund Å., 1975, A&A, 42, 407 Mashonkina L., Gehren T. 2001, A&A, 376, 232 Matteucci F., Raiteri C.M., Busso M., Gallino R., Gratton R. 1993, A&A, 272, 421 McClure R.D. 1983, ApJ, 268, 264 McClure R.D. 1984, PASP, 96, 117 McClure R.D., Fletcher J.M., Nemec J. 1980, ApJ, 238, L35 McWilliam A. 1997, ARA&A, 35, 503 McWilliam A., 1998, AJ, 115, 1640 McWilliam A., Preston G.W., Sneden C., Searle L. 1995, AJ, 109, 2757 McWilliam A., Rich R.M. 1994, ApJS, 91, 749 Mel´endez J., Barbuy B., 2006, in preparation Mel´endez J., Ramirez I. 2003, A&A 398, 705 Mennessier M.O., Luri X., Figueras F., G´omez A.E., Grenier S., Torra J., North P. 1997, A&A 326, 722 Meyer B.S., 1994, ARA&A, 32, 153 Mishenina T.V., Kovtyukh V.V., Soubiran C., Travaglio C., Busso M. 2002, A&A, 396, 189 Nissen P.E., Chen Y.Q., Schuster, W.J., Zhao G. 2000, A&A, 353, 722 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars 31 Nissen P.E., Hφg E., Schuster W. 1997, in Hipparcos, Venice Vergely J.L., Freire Ferrero R., Egret D., Koppen J. 1998, '97(ESO SP-402, p. 225) A&A, 340, 543 Nomoto K., Koichi I., Nobuhiro K. 1997b, Science, 276, 1378 North P., Berthet S., Lanz T. 1994, A&A 281, 775 Pagel B.E.J. 1968, em Origin and Distribution of the Elements, Warner, B. 1965, MNRAS, 129,263 Woolf V.M., Tomkin J., Lambert D.L. 1995, ApJ, 453, 660 Woosley S.E., Hartmann D., Hoffman R.D., Haxton W. 1990, ed. LH Ahrens, pp. 195-204, Oxford: Pergamon ApJ, 356, 272 Palmeri P., Quinet P., Wyart J.F., Bi´emont E. 2000, PhyS, 61, 323 Woosley S.E., Weaver T.A. 1995, ApJS, 101, 181 Yong D., Grundahl F., Lambert D.L., Nissen P.E., Shetrone Pereira C.B. 2005, AJ, 129, 2469 Pereira C.B., Junqueira S. 2003, A&A, 402, 1061 Perryman M. A. C., Lindegren L., Kovalevsky J. et al., 1997, M.D. 2003, A&A, 402, 985 Zhao G., Magain P. 1990, A&A, 238, 242 Zoccali, M., Barbuy, B., Hill, V., et al. 2004, A&A, 423, 507 A&A, 323, L49 Peterson R.C., Kurucz R.L., Carney B.W. 1990, ApJ, 350, 173 Plez B., Brett J.M., Nordlund A. 1992, A&A, 256, 551 Przybilla, N., Butler, K., Kudritzki, R.-P. 2001, A&A, 379, 936 Prochaska J.X., Naumov S.O., Carney B.W., McWilliam A., Wolfe A.M. 2000, AJ, 120, 2513 Ram´ırez I., Mel´endez J. 2004, A&A 417, 301 Reddy B.E., Tomkin J., Lambert D.L., Allende Prieto C. 2003, MNRAS, 340, 304 Rossi S., Beers T. C.; Sneden C., Sevastyanenko T., Rhee J., Marsteller B., 2005, AJ, 130, 2804 Rutten R.J., 1978, SoPh, 56, 237 Ryan S.G., Norris J.E., Beers T.C. 1996, ApJ, 471, 254 Sackmann I.-J., Boothroyd A.I. 1992, ApJ, 392, L71 Sackmann I.-J., Boothroyd A.I. 1999, ApJ, 510, 217 Simmerer J., Sneden C., Cowan J.J., Collier J., Woolf V.M. Lawler J.E. 2004, ApJ, 617, 1091 Simmerer J., Sneden C., Ivans I., Kraft R.P., Shetrone M.D., Smith V.V. 2003, ApJ, 125, 2018 Sivarani, T., Bonifacio, P., Molaro, P. et al. 2004, A&A, 413, 1073 Smith V.V., Coleman H., Lambert D.L. 1993, ApJ, 417, 287 Smith V.V., Lambert D.L. 1986, ApJ, 303, 226 Smith V.V., Suntzeff N.B., Cunha K., Gallino R., Busso M., Lambert D.L. Straniero O. 2000, AJ, 119, 1239 Sneden C., Crocker D.A. 1988, ApJ, 335, 406 Sneden C., Gratton R.G., Crocker D.A. 1991, A&A, 246, 354 Sneden C., McWilliam A., Preston G.W., et al., 1996, ApJ, 467, 819 Spite, F., Spite, M. 1975, A&A, 40, 141 Spite M., Spite F. 1978, A&A, 67, 23 Spite M., Spite F. 1982, Nature, 297, 483 Steffen M. 1985, A&AS, 59, 403 Th´evenin F., 1989, A&AS, 77, 137 Th´evenin F., 1990, A&AS, 82, 179 Th´evenin F., Idiart T.P., 1999, ApJ, 521, 753 Th´evenin F., Jasniewicz G. 1997, A&A, 320, 913 Timmes F.X., Woosley S.E., Weaver T.E. 1995, ApJS, 98, 617 Tomkin J., Lambert D.L. 1999, ApJ, 523, 234 Tomkin J., Lambert D.L., Edvardsson B., Gustafsson B., Nissen P.E. 1989, A&A 219, L15 Udry S., Jorissen A., Mayor M., Van Eck S. 1998a, A&AS, 131, 25 Udry S., Mayor M., Van Eck S., Jorissen A., Pr´evot L., Grenier S., Lindgren H. 1998b, A&AS, 131, 43 Van Eck S., Goriely S., Jorissen A., Plez B. 2003, A&A, 404, 291 32 D.M. Allen and B. Barbuy: Analysis of 26 Barium Stars Table 19. Abundance uncertainties. HD 150862 HD 5424 X σlogǫ(X)⊙ 0.10 Li 0.06 C 0.06 N 0.06 O Na 0.03 0.05 Mg 0.07 Al 0.05 Si 0.02 Ca 0.10 Sc Ti 0.06 0.02 V 0.03 Cr 0.04 Co 0.04 Ni 0.04 Cu Zn 0.08 0.07 Ss 0.07 Sr 0.03 Y 0.02 Zz 0.02 Zr 0.05 Mo Ba 0.05 0.03 La 0.04 Ce 0.15 Pr Nd 0.01 Sm 0.06 Eu 0.01 0.04 Gd 0.06 Dy Pb 0.08 0.13 0.08 0.13 0.10 0.05 0.05 0.06 0.04 0.05 0.06 0.05 0.06 0.05 0.05 0.04 0.09 0.05 0.09 0.15 0.04 0.06 0.05 0.13 0.05 0.06 0.05 0.06 0.06 0.08 0.10 0.10 0.06 0.19 ... 0.09 0.13 0.11 0.04 0.06 0.08 0.05 0.03 0.11 0.07 0.05 0.04 0.05 0.04 0.09 0.09 0.11 0.16 0.04 0.04 0.04 0.13 0.06 0.05 0.05 0.16 0.05 0.09 0.09 0.10 0.07 0.20 n log Ap log AT log Am log Alg log Av σlog ǫ(X) σ[X/Fe] n log Ap log AT log Am log Alg log Av σlog ǫ(X) σ[X/Fe] 1 1.10 1 8.96 1 8.98 2 8.97 6 6.41 3 7.55 2 6.32 5 7.36 6 6.26 4 3.42 12 5.02 7 3.81 6 5.62 4 4.99 10 6.30 2 4.01 4 4.60 2 3.87 1 3.67 12 3.30 4 3.50 5 3.75 1 2.42 5 3.18 8 1.96 10 2.28 2 1.06 8 2.04 4 1.26 3 0.82 2 1.17 2 1.34 1 2.65 2 1.05 4 8.90 12 8.26 2 8.92 6 6.53 3 7.80 2 6.57 5 7.64 6 6.11 4 3.57 12 4.87 13 3.91 6 5.72 5 5.27 10 6.50 3 4.06 4 4.60 3 3.92 2 3.57 12 3.54 5 3.30 5 3.90 1 2.12 5 3.73 8 2.67 11 3.08 3 2.16 9 3.14 5 2.26 4 1.19 3 1.67 2 2.94 1 3.05 0.23 0.20 0.18 0.24 0.20 0.20 0.24 0.20 0.25 0.22 0.20 0.19 0.19 0.21 0.19 0.21 0.21 0.21 0.24 0.19 0.20 0.20 0.25 0.19 0.19 0.19 0.25 0.20 0.20 0.21 0.38 0.28 0.29 1.00 8.93 8.28 9.07 6.53 7.80 6.67 7.69 6.26 3.72 4.82 3.86 5.77 5.42 6.55 4.11 4.70 3.92 3.67 3.64 3.20 4.00 2.17 3.78 2.77 3.23 2.36 3.04 2.36 1.29 1.97 2.79 3.00 1.05 8.90 8.26 8.92 6.48 7.78 6.62 7.59 6.36 3.62 4.77 3.91 5.77 5.37 6.55 4.01 4.65 3.93 3.47 3.53 3.15 3.80 2.14 3.78 2.67 3.09 2.15 2.94 2.21 1.14 1.87 2.79 3.00 1.10 8.96 8.98 8.87 6.41 7.55 6.32 7.36 6.26 3.42 5.02 3.86 5.62 5.02 6.25 4.01 4.60 3.87 3.67 3.30 3.45 3.75 2.42 3.23 1.96 2.33 1.06 2.09 1.31 0.92 1.07 1.34 2.65 1.10 8.96 8.98 8.97 6.41 7.55 6.32 7.36 6.26 3.42 5.02 3.81 5.62 4.99 6.30 4.01 4.60 3.87 3.67 3.30 3.50 3.75 2.42 3.18 1.96 2.28 1.06 2.04 1.26 0.82 1.17 1.34 2.65 0.90 8.92 8.21 8.97 6.48 7.75 6.47 7.59 6.21 3.62 4.62 3.61 5.67 5.32 6.45 4.01 4.70 3.82 3.47 3.49 3.30 3.90 2.02 3.68 2.62 3.13 2.21 3.04 2.16 1.21 1.87 2.69 2.85 0.86 8.72 8.11 8.78 6.29 7.61 6.38 7.45 5.97 3.41 4.65 3.72 5.58 5.13 6.36 3.87 4.46 3.73 3.38 3.35 3.11 3.71 1.95 3.63 2.53 2.95 1.97 2.95 2.15 1.00 1.48 2.75 2.86 1.20 9.01 9.08 8.87 6.46 7.60 6.37 7.38 6.31 3.32 4.92 3.91 5.67 5.02 6.29 4.11 4.65 3.97 3.79 3.35 3.55 3.80 2.52 3.23 2.06 2.33 1.11 2.14 1.36 0.92 1.07 1.29 2.80 1.06 8.92 8.94 8.93 6.37 7.51 6.28 7.32 6.22 3.38 4.98 3.77 5.58 4.95 6.26 3.97 4.56 3.83 3.63 3.26 3.46 3.71 2.38 3.14 1.92 2.24 1.02 2.00 1.22 0.78 1.13 1.30 2.61 ... 0.10 0.07 0.17 0.09 0.11 0.17 0.09 0.17 0.15 0.10 0.08 0.07 0.12 0.06 0.11 0.13 0.12 0.18 0.06 0.10 0.10 0.18 0.08 0.07 0.08 0.23 0.09 0.10 0.10 0.34 0.22 0.25 Table 20. Abundances found in the literature for barium stars. The stars codes in the header are the same as in Table 12. m=average of Zr  and Zr . References: E93 - Edvardsson et al. (1993); T89 - Tomkin et al. (1989); P05 - Pereira (2005); B92 - Barbuy et al. (1992); L03 - Liang et al. (2003); N94 - North et al. (1994); P03 - Pereira & Junqueira (2003); S86 - Smith & Lambert (1986); L91 - Lu (1991); S93 - Smith et al. (1993). HD/HR e2 E93 T89 ... Li ... C ... N ... O 0.13 Na Mg 0.20 Al 0.06 Si 0.12 0.05 Ca ... Sc 0.04 Ti ... V ... Cr Co ... Ni 0.12 ... Cu ... Zn ... Sr  ... Sr  Y 0.50 Zr  ... Zr  0.65 ... Mo 0.54 Ba ... La ... Ce ... Pr ... Nd Sm ... ... Eu ... 0.1 0.0 0.1 ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 0.7 ... 0.4 ... 0.4 ... 0.3 ... ... ... 0.2 e4 P05 ... 0.71 ... ... 0.08 0.07 ... 0.32 0.12 -0.15 ... ... 0.07 ... -0.05 -0.07 0.04 ... ... 0.75 ... 0.71 ... 1.17 0.75 0.68 ... 0.80 ... 0.19 e6 P05 ... 0.46 ... ... 0.21 -0.09 ... 0.39 0.03 0.12 ... ... ... ... 0.07 0.03 0.04 ... ... 0.09 ... ... ... 1.38 1.23 0.92 ... 0.53 ... ... e7 e8 e9 P05 B92 L03 N94 ... 0.64 ... ... 0.02 -0.03 -0.13 0.24 0.04 -0.06 ... ... 0.12 ... -0.02 0.06 0.06 ... ... 0.72 ... ... ... 0.75 0.56 0.29 ... 0.07 ... 0.26 ... 0.15 0.70 0.05 ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 0.39 0.16 0.10 0.03 0.21 -0.01 0.00 -0.17 -0.33 0.26 ... -0.01 ... ... ... ... 0.47 0.41 ... ... 0.67 0.81 ... ... ... ... 0.36 ... 0.70 ... 0.76 ... ... 0.29 0.30 0.41 ... 0.11 ... ... ... -0.18 ... ... ... ... 0.70 ... 0.88 ... ... ... ... ... 1.22 ... ... e10 N94 ... 0.68 ... 0.76 ... ... ... ... 0.20 ... -0.10 ... -0.11 0.07 -0.04 ... ... ... ... 1.21 ... 0.90 ... 1.44 ... 1.02 ... 0.87 ... ... e11 P03 ... 0.61 ... ... 0.00 0.10 0.14 0.21 0.07 -0.13 0.25 -0.08 0.21 0.25 0.04 0.01 0.26 ... ... 1.01 1.22 ... ... 1.51 1.75 1.32 ... 0.97 ... 0.61 e12 S86 L91 S93 <1.3 ... ... ... ... ... 0.02 ... 0.11 ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 0.47 ... ... 0.23 -0.24 -0.23 0.15 0.20 0.13 0.29 0.19 0.16 0.39 -0.03 ... -0.23 ... ... 0.86 0.62 ... ... 0.87 0.78 0.57 1.01 0.60 ... ... ... 0.61 ... 0.27 0.07 ... ... 0.16 0.15 ... ... -0.07 0.17 0.09 0.15 0.08 ... ... ... 1.11 ... 0.89m ... 0.83 ... ... ... 0.60 ... ... e13 N94 ... 0.26 ... 0.37 ... ... 0.05 ... 0.38 ... -0.03 ... 0.01 0.13 0.02 ... ... 0.66 ... 0.75 ... 0.52 ... 0.85 ... 0.77 ... 0.63 ... ... e14 N94 ... 0.40 ... 0.40 ... ... -0.02 ... 0.52 ... -0.13 ... -0.04 0.29 0.09 ... ... 0.96 ... 1.09 ... 0.68 ... 0.77 ... 0.54 ... 0.17 ... ... e15 N94 ... 0.77 ... 1.15 ... ... ... 0.48 0.37 ... 0.12 ... -0.07 0.06 0.06 ... ... 1.09 ... 1.29 ... 0.95 ... 1.54 ... 1.43 ... 1.15 ... ... e18 e19 e20 N94 L91 N94 N94 L91 ... 0.87 ... ... 0.21 0.25 0.52 0.09 0.04 -0.08 0.37 ... 0.29 0.69 -0.01 ... ... ... ... 1.04 ... ... 1.12 1.32 0.96 1.42 ... 0.98 1.00 ... 0.75 ... 0.57 ... ... ... ... 0.03 ... -0.14 ... -0.32 -0.03 -0.21 ... ... 1.14 ... 1.09 ... 0.89 ... 1.40 ... 1.35 ... 1.18 ... ... ... 0.53 ... 0.73 ... ... 0.36 0.33 0.54 ... ... ... ... ... 0.10 ... ... ... ... 1.32 ... 1.13 ... ... ... ... ... 1.19 ... ... ... ... ... ... 0.17 0.49 ... ... 0.25 -0.30 -0.12 0.13 0.11 -0.15 0.06 ... -0.25 ... ... 0.65 ... 0.39 ... ... 0.38 0.35 ... 0.32 0.18 ... 0.42 ... 0.58 ... ... -0.04 0.28 0.47 ... -0.05 ... 0.30 0.57 0.11 ... ... 1.23 ... 1.26 ... 0.90 ... 0.97 ... 0.69 ... 0.55 ... ... e21 N94 ... 0.36 ... 0.34 ... ... -0.13 0.15 0.23 ... -0.28 ... -0.22 0.03 -0.09 ... ... 0.89 ... 0.93 ... 0.83 ... 1.25 ... 0.90 ... 0.87 ... ...
astro-ph/0406372
1
0406
2004-06-16T18:40:17
Linear line polarimetry modelling of pre-main sequence stars
[ "astro-ph" ]
We present emission line polarimetry data and modelling relevant to the circumstellar geometry and kinematics around pre-main sequence stars. For a sample of both Herbig Ae/Be stars and T Tauri stars, we find that most show polarization changes across Halpha, implying that flattened structures are common on the smallest scales -- and over a range of stellar masses. We also present Monte Carlo calculations of spectral line profiles scattered in rotating accretion disks. We consider both the case of a central star that emits line photons uniformly, as well as via hot spots. Intriguingly, the switch between a uniform point source and a finite-sized star results in a marked difference in the position angle variation across the line. Our models demonstrate the diagnostic potential of line polarimetry in determining the disk inclination and the size of the inner hole -- a spatial scale no other technique currently accesses.
astro-ph
astro-ph
Linear line polarimetry modelling of pre-main sequence stars Jorick S. Vink1, Janet E. Drew1, Tim J. Harries2, Rene D. Oudmaijer3 1 Imperial College London, Physics Department, Prince Consort Road, London SW7 2AZ, UK 2 University of Exeter, School of Physics, Stocker Road, Exeter EX4 4QL, UK 3 University of Leeds, School of Physics & Astronomy, EC Stoner Building, Leeds LS2 9JT, UK Abstract. We present emission line polarimetry data and modelling relevant to the circumstellar geometry and kinematics around pre-main sequence stars. For a sample of both Herbig Ae/Be stars and T Tauri stars, we find that most show polarization changes across Hα, implying that flattened structures are common on the smallest scales -- and over a range of stellar masses. We also present Monte Carlo calculations of spectral line profiles scattered in rotating accretion disks. We consider both the case of a central star that emits line photons uniformly, as well as via hot spots. Intriguingly, the switch between a uniform point source and a finite-sized star results in a marked difference in the position angle variation across the line. Our models demonstrate the diagnostic potential of line polarimetry in determining the disk inclination and the size of the inner hole -- a spatial scale no other technique currently accesses. 1. General Introduction It is believed that low-mass stars form through the collapse of an interstellar cloud. During the subsequent pre-main sequence (PMS) T Tauri phase material is accreting from the disk onto the star, most likely through magnetospheric funnels (e.g. Johns-Krull et al. 1999). Whilst this basic picture of star formation is relatively well understood, problems relating to a star's angular momentum remain, as we have little information on the size of the disk inner hole. For intermediate mass (2 -- 10 M⊙) Herbig Ae/Be stars our knowledge becomes even more patchy, and for stars above 10 M⊙ there is not even any consensus on the mode of star formation itself. Traditionally, the switch between low-mass and high-mass star formation has been thought to occur at the T Tauri/Herbig boundary (at ≃ 2 M⊙), since this is where low-mass T Tauri stars possess convective envelopes, whilst in- termediate mass Herbig Ae star envelopes are radiative. However, recent data indicate that such a division is no longer tenable: Herbig Ae stars and T Tauri stars have a range of characteristics in common, varying from the presence of inverse P Cygni profiles, indicative of active accretion (Catala et al. 1999), to the detections of line polarizations, signalling rotating accretion disks (Vink et al. 2002, 2003). It is clear that what is needed to understand star formation as a function of mass (and ultimately understand the IMF) is observations of the 1 2 Jorick S. Vink et al near-star environment over a wide range of young stellar objects. Polarimetry across emission lines is just such a tool. Triplots of the observed polarization spectra of (a) the T Tauri Figure 1. star RY Tau (Vink et al. 2003), the Herbig Ae star XY Per (Vink et al. 2002), and the Herbig Be star HD 53367 (Oudmaijer & Drew 1999). In all plots, the Stokes I spectrum is shown in the lowest panel, the %Pol is indicated in the middle panel, whilst the position angle, θ, is plotted in the upper panel. 2. Line polarimetry Spectropolarimetry is a powerful tool to study the near-star regions of PMS stars and to determine their geometries. The technique has widely been applied to Line polarimetry of PMS stars 3 Figure 2. QU representations of the observed polarization spectra of the T Tauri star RY Tau and the Herbig Ae star MWC 480. The arrow denotes the sense of increasing wavelength. Note the resemblance in the loops between the Herbig Ae and T Tauri star in the QU diagram. early-type stars, where circumstellar free electrons -- e.g. in a disk -- are able to polarize the continuum light more than the line photons. This is widely known as 'depolarization' (see Fig. 1(c) and Oudmaijer; these proceedings). In this case, the polarization angle (PA) of the polarization does not change across the line (if it is corrected for foreground), and the shape of the phenomenon in the QU plane simply involves a straight line (independent of foreground). In certain circumstances it is feasible that the line photons are scattered and polarized themselves (e.g. McLean 1979). Wood et al. (1993) performed an analytical study of the polarization and PA of a uniform point source that is scattered within a surrounding moving medium. Specifically, for a rotating disk, they found that the PA is no longer constant through the line, but rotates by a few degrees, resulting in a 'loop' in the QU plane (see Fig. 2 for examples), which they attributed to stellar occultation. The diagnostic value of this PA flip (QU loop) is that it is the direct signpost for the presence of rotation. 3. Data of T Tauri and Herbig Ae/Be stars In recent years, we have surveyed T Tauri and Herbig Ae/Be stars spectropo- larimetrically. For Herbig Be stars, the frequency of depolarizations was found to be essentially the same as was found for classical Be stars in the 1970s -- indicating they are embedded in electron scattering disks (see Fig. 1(c) for the Herbig Be star HD 53367 and Oudmaijer, these proceedings). When observing later spectral type PMS stars, one might perhaps expect to witness a sharp decrease in the number of polarization line effects, because the amount of free electrons is anticipated to drop. Furthermore, narrow-band filter work in the 1980s indicated a general absence of polarization changes across Hα in T Tauri stars (e.g. Bastien 1982). However, this is not the case when 4 Jorick S. Vink et al Figure 3. Monte Carlo predictions of the line polarization for the cases of a finite-sized line-emitting star embedded in a scattering disk with (left) and without (right) an inner hole. The disk is inclined at 45o and has an inner hole of 5 times the stellar radius (in the left panel). observing with higher spectral resolution. Typical polarization spectra with a resolution of R ≃ 9000 of the T Tauri star RY Tau and the Herbig Ae star XY Per are presented in Figs. 1(a) and (b). The S-shaped PA flips in these data are indicative of line polarization, and the resulting loops in the QU plane (Fig. 2) show that a rotating disk geometry is the dominant factor in producing the line effects in the majority of Herbig Ae (9/11; Vink et al. 2002) and T Tauri stars (9/10; Vink et al. 2004, in preparation). 4. Line polarization models Given the common occurrence of these QU loops in Herbig Ae and T Tauri stars, we are developing polarization models of line emission scattered off rotating disks. In particular, we employ the 3D Monte Carlo model torus (Harries 2000) to simulate both uniform as well as asymmetric illumination (by two diametrically opposed hot spots) onto a rotating scattering disk with, or without, a significant inner hole. These options of truncating the inner disk, and studying illumination via hot spots, are motivated by the growing interest in the magnetic accretion models for both T Tauri (e.g. Edwards et al. 1994), and Herbig Ae stars (e.g. Vink et al. 2002). Figure 3 shows an intriguing result for the case of a uniformly radiating star. We find that there is a marked difference between scattering of line emission by a disk that reaches the stellar surface (Fig. 3b), and a disk with a significant inner hole (Fig. 3a). The single position-angle flip, seen on the left-hand side is similar to that predicted by the analytic models of Wood et al. (1993) -- but the double Line polarimetry of PMS stars 5 PA flip as seen on the right-hand side, associated with the undisrupted disk, is a surprise. This effect is due to the geometrically correct treatment of the finite- sized star interacting with the disk's rotational velocity field (Vink et al. 2004). Since a gradual increase of the hole size transforms the double rotations smoothly back into single ones -- as the line emission object approaches that of a point source -- our models demonstrate the diagnostic potential of line polarimetry in determining the disk inclination and the size of the inner hole. By changing the configuration to a non-uniformly line emitting object, such as one where the emission originates from hot spots, we find that the line po- larimetry depends strongly on the rotational phase. Stassun & Wood (1999) have shown that the magnetic accretion model can account for the observed periodic changes in the PA and polarization of continuum light. However, pho- topolarimetry does not provide diagnostics of the disk truncation radius. We have therefore extended Stassun & Wood type-models to spectral lines. First results indicate that there are significant changes in the shapes and amplitudes of the PA and polarization across the spectral line as a function of rotational phase. 5. Summary We have presented data and modelling results of line polarimetry for PMS stars. For the Herbig Ae/Be stars, we found that a large majority show a line effect -- indicating flattened circumstellar geometries. Interestingly, we found a marked difference between the Herbig Be stars and groups of later spectral type. For the Herbig Be stars, electron scattering disks can explain the depolarisations. At lower masses, more complex behaviour appears across Hα. Here the concept of compact line emission scattered off a rotating disk may explain the observed QU loops. We have also presented polarimetric line profiles for scattering off accretion disks calculated with a Monte Carlo code. We considered the cases of a central object that emits line photons (a) uniformly, and (b) via hot spots only. For case (a), the switch between a point source and a finite-star photon source results in a surprising difference in the shapes of the predicted position angles. Most notably, we find double PA rotations. For case (b), emission from stellar hot spots, we find polarization signatures that are strongly dependent on hot spot phase with respect to the observer. Rotational modulation of line polarization shows great promise for unravelling the complexities occurring in the circumstellar environments around young stars. References Catala C., Donati J. F. Bohm T., et al., 1999, A&A 345, 884 Harries T.J., 2000, MNRAS 315, 722 Johns-Krull C.M., Valenti J.A., Koresko C., 1999, ApJ 516, 900 Oudmaijer R.D., & Drew J.E., 1999, MNRAS 305, 166 Stassun K., & Wood K., 1999, ApJ 510, 892 Vink J.S., Drew J.E., Harries T.J., Oudmaijer R.D., 2002, MNRAS 337, 356 Vink J.S., Drew J.E., Harries T.J., Oudmaijer R.D., 2003, A&A 406, 307 Vink J.S., Harries T.J., Drew J.E., 2004, A&A, submitted
astro-ph/0011252
1
0011
2000-11-13T19:04:52
Concluding Remarks on New Cosmological Data and the Values of the Fundamental Parameters
[ "astro-ph" ]
I review the reason for considering the prime purpose of the program of measurements of the fundamental parameters of cosmology to be the tests of cosmological models. I comment on the philosophy by which we are approaching this goal, offer an assessment of where we stand, and present some thoughts on where the tests may be headed.
astro-ph
astro-ph
New Cosmological Data and the Values of the Fundamental Parameters IAU Symposium, Vol. 201, 2000 A. N. Lasenby, A. W. Jones and A. Wilkinson Concluding Remarks P. J. E. Peebles Joseph Henry Laboratories, Princeton University, Princeton, NJ 08544, USA I review the reason for considering the prime purpose of the Abstract. program of measurements of the fundamental parameters of cosmology to be the tests of cosmological models. I comment on the philosophy by which we are approaching this goal, offer an assessment of where we stand, and present some thoughts on where the tests may be headed. 1. Introduction These Proceedings document impressive progress toward a satisfactory comple- tion of the great program of measurements of the parameters of cosmology that commenced in the 1930s. It has taken a long time, and has brought into play many phenomena and measurements that could not have been anticipated in the 1930s. We may at last be approaching closure of this program, and it is appropriate to reflect on why we are so interested in these measurements and how it informs our interpretation of the results. 2. The Significance of the Cosmological Tests I take the literal reading, that the purpose of the cosmological tests is to test models, in particular the commonly accepted relativistic Friedmann-Lemaıtre cosmology. It certainly is useful to have byproducts such as the demonstration of the presence of a term in the stress-energy tensor that acts like Einstein's cosmological constant Λ, which may help guide us to a resolution of the per- plexing physics of the energy density of the vacuum, and a measurement of the radius of curvature of space sections at constant world time, which may prove to be a clue to what the universe was like before it could have been described by the Friedmann-Lemaıtre model. But all this is true only if we have convincing reason to trust the basis for these results. A part of cosmology we can trust is the near homogeneous evolution of the observable universe from a denser hotter state. The list of evidence is familiar but worth repeating to make the point: we have compelling reason to believe this is what happened. Deep counts of objects at wavelengths ranging from radio to gamma rays are close to isotropically distributed across the sky. Either we are close to a center of spherical symmetry or our universe is close to homogeneous. If the latter, and the distribution is expanding so as to preserve homogeneity and isotropy, the recession velocity satisfies Hubble's law. The low redshift part of the SNeIa measurements is an impressively tight demonstration of the 1 2 Peebles redshift-distance relation. The cosmological interpretation of quasar redshifts passes demanding tests, such as the tight correlation of Lyman-limit and Mg II absorption lines with galaxies at close to the same angular position and redshift, showing quasars are behind lower redshift galaxies. If the expansion traces back to very high density galaxies at high redshift are seen as they were closer to the time when galaxies could not have existed, and ought to look younger than nearby ones. The effect is amply demonstrated. The 3 K radiation (the CBR) could not have relaxed to its thermal spectrum in the universe as it is now because space is not opaque at the Hubble length: radio sources are observed at z ∼ 1. We can understand the thermal spectrum if the universe has expanded from a denser, hotter state that is optically thick within the Hubble length. The angular position of the peak of the spectrum of angular fluctuations of the CBR agrees with the conventional physics of the evolution of primeval adiabatic mass density fluctuations if the universe has expanded and cooled by a factor much larger than zeq ∼ 1000, the redshift of decoupling of matter and radiation. Helium and deuterium are natural byproducts of expansion from still higher temperature. This list offers no guidance to what happened at very large redshift, or well outside the Hubble length. The inflation concept has shown us how easy it is to imagine the universe at great distance is not at all like what we see, but determining whether such an "island universe" picture is realistic is outside the current round of cosmological tests as I would define them. I don't know whether the people in the 1930s who pioneered the program of cosmological tests gave much thought to the spatial and temporal limitations of empirical evidence within their cosmology. If not we have to adjust the program. You can add to the list of evidence for evolution, depending on how much you want to rely on models, but I think the point is clear: it employs a broad variety of phenomena observed in quite different ways. Individual entries could be wrong, but it would be absurd to imagine all quite consistently point in the wrong direction. Thus Hoyle, Burbidge & Narlikar (1993) accept cosmic evolu- tion, but argue the last substantial addition to the entropy in the CBR could have occurred at a much more modest expansion factor than in the standard model. This is a considerable difference, but it should not obscure the point that the redundancy of evidence has forced us to the answer to Hubble's (1936) question, is the cosmological redshift the result of the general recession of the nebulae? It is, and the general recession is associated with cosmic evolution. Our answer to Hubble depends on local physics and symmetry arguments, but it makes little use of general relativity theory (hereinafter GR); I did not even mention the relativistic relations among observables that Tolman (1934) listed.1 The observations of supernovae of type Ia probe one of the relations, between magnitude and redshift, and detect a departure from the Einstein- de Sitter case. This magnificent accomplishment is in no way depreciated by noting that by itself it is not a cosmological test. In addition to the slight 1The theory of the origin of the light elements assumes the expansion rate equation, H 2 = 8πGρ/3. This follows from local physics; relativity enters only in the expression for active gravitational mass. The CBR anisotropy computation uses the angular size distance-redshift relation, a highly nontrivial application of the large-scale spacetime geometry, but the success of the prediction forces its inclusion in my list of elementary evidence for cosmic evolution. New Cosmological Data 3 chance some quirk of the physics of supernovae has avoided the thorough checks for systematic error (or, to be really cautious, that Nature has put us at the center of a spherically symmetric universe with a slight radial density gradient), the conventional interpretation depends on GR, and, within this theory, the measurement is readily fitted by the adjustment of free parameters. Hoyle, Burbidge & Narlikar (1993) might similarly fit the measurement by suitable choice of parameters within their theory. The elegant logic of general relativity theory, and its precision tests, rec- ommend GR as the first choice for a working model for cosmology. But the Hubble length is fifteen orders of magnitude larger than the length scale of the precision tests, at the astronomical unit and smaller, a spectacular extrapola- tion. The extrapolation is tested by checking for consistency of the cosmological parameters derived from different aspects of the geometry of spacetime. The Robertson-Walker line element figures in Tolman's (1934) list of cosmological relations. The computation of the CBR anisotropy spectrum uses GR to prop- agate the irregularities in the radiation distribution through spacetime that is predicted to be strongly curved over the expansion factor z ∼ 1000 since de- coupling, and it uses GR to predict the dynamics of small fluctuations in the distributions of matter and radiation at z ∼ 1000. The dynamical estimates of galaxy masses from rotation curves and streaming velocities assume the latter aspect of GR, the inverse square force law for gravity. Weak and strong grav- itational lensing use this law, with the usual factor of two correction. A tight check of consistency of the parameters derived from these different phenomena would be a demanding test of GR and the cosmology. The spectrum of angular fluctuations of the CBR offers a wonderfully rich basis for these tests. These Proceedings discuss the constraints on the density parameters in dark matter and in Einstein's cosmological constant Λ (or a term in the stress-energy tensor that acts like Λ), to be compared to what is indicated by the dynamical measurements of masses of galaxies and systems of galaxies, by the curvature of the redshift-magnitude relation, and by the measurement of Hoto; the density parameter in baryons, to be compared to the theory and obser- vational tests of the origin of helium and deuterium at high redshift and to the observational baryon budget at low redshift; the density parameter in neutrinos, to be compared to laboratory and atmospheric oscillation experiments; and the amplitude of the primeval density fluctuations, to be compared to measurements of the distributions of galaxies and mass at low redshift. The impressive consistency of constraints that have already emerged from such different applications of GR and the cosmological principle suggests the theory and the cosmology are on the right track. We should be cautious about the details, however, because the interpretation of the CBR anisotropy also assumes the adiabatic cold dark matter (CDM) theory for structure formation, and the tests of this model depend on some subtle issues of astronomy. 3. The Model for Structure Formation and the Issue of Voids We pay particular attention to simple and elegant ideas in physical science be- cause Nature tends to agree with us. We have examples in cosmology: GR, Einstein's cosmological principle, and the adiabatic CDM model for structure 4 Peebles formation.2 But Nature is quite capable of surprising us, as witness the evidence for a significant cosmological constant, which a few years ago was generally con- sidered to have no socially redeeming value. Since many of the cosmological tests depend on the CDM model we must consider its empirical tests. In these Proceedings Carlos Frenk and John Peacock present impressive observational successes of the CDM model. There are a few clouds on the small- scale part of the horizon, however; an example that has particularly impressed me is the void phenomenon (Peebles 1989). Carignan and Freeman's (1988) "dark galaxy," DDO 154, seems to be a close approximation to one of the failed galaxies that figure in commonly discussed interpretations of numerical simu- lations of the CDM model. In these simulations there is appreciable mass in the voids defined by the positions of dark mass concentrations that are massive enough to qualify as homes for normal L ∼ L∗ high surface brightness galaxies. This void medium contains low mass halos that would seem to be acceptable homes for galaxies like DDO 154. So why are galaxies like DDO 154 not found in the voids? There are void galaxies; nearby examples are the pair NGC 6946 and NGC 6503. The former is an Arp (1966) peculiar galaxy, but only because a supernova was seen in it. Sandage and Bedke (1988) give a magnificent image of this galaxy; I have been assured it looks like an ordinary large near face-on spiral, though maybe unusually gas-rich. The other appears to be an edge-on spiral; it is classified as Scd in the Nearby Galaxies Catalog (Tully 1988). The CDM model simulations show occasional substantial upward mass fluctuations in generally low density void regions that could be homes for L ∼ L∗ void galax- ies, but how would the baryons in these isolated mass peaks get spun up to form normal-looking if isolated spirals? To me the most remarkable and challenging phenomenon is that observable objects respect the same voids. This applies to giant and dwarf galaxies, and low and high surface brightness ones (eg. Pustil'nik et al. 1995; Popescu, Hopp & Rosa 1999; and references therein); to gas clouds observed in emission (eg. Zwann et al. 1997); and to high surface density gas clouds observed in absorption (Lanzetta et al. 1995; Steidel, Dickinson & Persson 1994).3 A common and defensible opinion is that the astrophysics by which void matter becomes visible as a galaxy of stars or an HI or MgII absorber is so complicated as to quite confuse the interpretation of void phenomena. Cen & Ostriker (2000) give an example: in their physically motivated prescription for galaxy formation the void probability for all galaxies identified in a simulation is much larger than for the mass. Cen & Ostriker conclude observed voids are not an argument against CDM-like models. This is a valuable example of the subtlety of the astrophysics, but I am even more impressed by the presence of dwarf galaxies on the outskirts of the Local Group, isolated enough to seem 2One can think of many other arguably less elegant models for structure formation; you can trace through astro-ph my list of alternatives, each killed by the inexorable advance of the measurements. 3Shull, Stocke & Penton (1996) show that gas clouds detected as very low surface density Lyman α absorbers avoid dense galaxy concentrations. My impression is that they also avoid the voids, but that is a subject of work in progress by Shull and colleagues. New Cosmological Data 5 to be primeval rather than products of physical processes operating within the large galaxies. These dwarfs are visible; why should similar primeval halos in the voids be so cunningly hidden? If void phenomena ruled out the CDM model we could turn to alternatives. Bode, Ostriker & Turok (2000) show that if the CDM is replaced by warm dark matter it greatly reduces the numbers of small dark mass halos, tends to produce dwarfs at lower redshift, and yields smooth patches of dark matter within the voids outlined by the massive halos. All are positive changes from CDM. But their figures 4 and 5 show caustics of dark matter threading the voids. If these caustics fragmented into low mass halos would the model predict greater numbers of dwarf or irregular galaxies extending into the voids than is observed? If the caustics remained smooth would the model predict more void absorption line systems than is observed? It looks like a serious challenge. I don't consider the void issue a very serious challenge to the Friedmann- Lemaıtre cosmology. Maybe I'm fooled by the astrophysics, as Cen & Ostriker (2000) argue. Maybe the CDM model must be adjusted, perhaps along the lines of Bode, Ostriker & Turok (2000), perhaps in some other way. The magnificent prediction of the measured first peak of the CBR angular fluctuation spectrum shows the CDM model very likely is close to the right picture. It would be less surprising to learn an improved structure formation model yields somewhat different constraints on the cosmological parameters, of course. Here is a worked example. Suppose at z ∼ 1000 there were objects with strong Lyman α emission lines, like quasar spectra with suppressed ionizing radiation; maybe primeval black holes. The Lyman α photons would delay recombination, preserving the height of the first peak of the CBR fluctuation spectrum, but shifting it to a larger angular scale for given cosmological parameters, and biasing this measure of space curvature (Peebles, Seager & Hu 2000). 4. Is Cosmology a Science? Disney (2000) asks whether cosmology "is a science at all," while I have been presenting it as a healthy and productive quantitative physical science. We might get some insight into the origin of these very different assessments from two considerations. First, our commonly accepted cosmology did grow by the introduction of hy- potheses to fit phenomena. Some hypotheses have been checked and established, as cosmic evolution. Some are being checked, as dark matter. The dynamical mass estimates quite consistently indicate the cosmological density parameter is low, Ωm ∼ 0.2. The SNe redshift-magnitude relation and the measurement of Hoto both favor this low value of Ωm. As discussed in the last section we are assuming GR, but applying it in quite different ways. The consistency at the level we now have is an elegant though not yet very precise test of GR and the dark matter hypothesis. In short, cosmology does depend on hypotheses, but we have nontrivial progress in testing them. Examples of work in progress are worth listing as a reminder of how broadly based the cosmological tests are becoming. Consider the projects to measure the predicted secondary peaks of the CBR temperature fluctuation spectrum and give us a first look at the polarization anisotropy, to test delayed recombination 6 Peebles among other things; measure the shape of the redshift-magnitude relation at redshifts well above unity, to test the prediction that the expansion becomes matter-dominated; establish the constraint on parameters from the rate of grav- itational lensing of background AGNs by galaxies; improve the measurement of Hoto to the point that it can distinguish between low density models with and without Λ; improve the constraints on the amount and distribution of mass from measurements of galaxy distributions, peculiar motions, and gravitational lens- ing; check the theory of structure formation through X-ray, optical, infrared, and radio surveys of the evolution of the intergalactic medium, galaxies, and clusters of galaxies; and maybe even test the dark and Λ-matter hypotheses through advances in particle physics. This work may yet lead us to an impasse, hypotheses multiplying faster than the data. That would drive us to new ideas, which would be exciting. The alternative is that we end up with an extensive and compellingly tight network of tests of a cosmology close to what we have now, which would be gratifying. The second consideration compares two lines of research. In his contribu- tion to these Proceedings Neil Turok considers what the universe might have been like at redshifts so high the Friedmann-Lemaıtre model certainly could not have applied. Turok very correctly emphasizes that the question is open and absolutely must be addressed. But we have to live with the fact that an empir- ical validation of the answer may be a long time coming. Most papers in these Proceedings deal with the more limited goal of understanding the large-scale nature of spacetime and its material content within our Hubble length, now and back in time through some ten orders of magnitude of expansion factor. This certainly is not a modest program either, but the empirical situation is remark- ably good: the standard theory has passed demanding observational tests, and work in progress promises substantial improvements. The empirical basis for research on the early universe is a lot more limited. This is an example of the ways in which the well tested and established cosmology is incomplete; another is that we can't say what the dark matter is. But any active physical science is similarly incomplete: each has a well-tested center around which is the exciting confusion of ongoing research. We all can make pretty good judgments about which elements of our subject are well and reliably established, and which are working hypotheses, when we put our minds to it. And the community has a reasonably accurate calibration of where each of us may tend to err on the side of caution or optimism. Our colleagues in other fields can't be expected to make these calls, and we shouldn't be surprised that when they have do so they may arrive at unduly pessimistic conclusions. We know how to remedy this, and should put our minds to it. I considered cosmology a real physical science decades ago, though with a meager well-established center. The big recent change has been the rate of addition to the established center. But I don't think we're in danger of running out of meaningful research on open issues any time soon. New Cosmological Data 7 5. A Next Generation of Cosmological Tests Martin Rees comments on the future of research in cosmology once the present round of tests is satisfactorily concluded. Here I add some thoughts on another round of cosmological tests of the physics of the very early universe. The rules of evidence in science have evolved to admit quite indirect ap- proaches. The community agrees that the many laboratory tests of quantum mechanics fully validate it as a real and magnificently successful physical sci- ence, even though no one has ever seen a state vector in nature. If the CBR revealed a distinctive signature of the tensor curvature fluctuations predicted in some implementations of inflation then I think most of us would accept it as an indirect but strong piece of evidence that inflation really did happen, even though none of us was there to see it. One version of the deconstructionist picture of science as I read about it is that clever people make up internally consistent stories to fit agreed-upon conditions, and that another group could have made up another story, equally consistent, with an equally satisfactory fit to some similar or maybe different set of agreed-upon conditions. Those of us who believe we have convincing evidence physical science describes aspects of an objectively real world, even on scales very different from what we can hold in our hands, reply that our theories have been validated by agreement with tightly over-constrained and cross checked empirical tests. Inflation as we now understand it can be adjusted to fit a broad range of possible empirical results. This situation is unnervingly close to the deconstructionist picture unless we stipulate that inflation is a working hypothesis. Michael Turner has asked whether empiricists like me would promote infla- tion from working hypothesis to established science if advances in basic physics produced a unified fundamental theory that is internally consistent, passes all laboratory tests, and predicts fields and interactions that unambiguously pro- duce inflation. If this fundamental theory allowed no free parameters to be adjusted to fit the astronomy, and within the uncertainties of the astrophysics it predicted the full suite of observations, it would be a brilliant addition to estab- lished cosmology. A less perfect fundamental theory might have free parameters, some of which could be fixed by laboratory measurements, while others would have to be determined by the constraints from a cosmology established through the rules of evidence one sees applied in these Proceedings. Should we be sat- isfied if this theory could be adjusted to fit all the observations? We would be well advised to adopt it in most of our analyses of astronomy, but not to accept an adjustment of the rules of evidence to admit it as the established picture. I suppose most of us think of the early universe as something that really hap- pened, and things that happen tend to leave traces. Let us cling to the hope that something will turn up. 6. Concluding Remarks The evidence assembled in these Proceedings favors the existence of several kinds of matter: one that acts like Einstein's cosmological constant, with density parameter ΩΛ ∼ 0.75, nonbaryonic low pressure matter with density parameter 8 Peebles ΩDM ∼ 0.2, baryons with density parameter Ωbaryons ∼ 0.05, and neutrinos with Ων ∼ 0.001. I believe it is too soon to add the first number to the list of firmly established elements of the science of cosmology, because it depends on the model for structure formation, and some of us see apparent problems with the model. People have been discussing the cosmological parameters for seven decades; we can wait a few more years to determine whether we have got the science right. If advances in the applications of the cosmological tests firmly established the values of the fundamental parameters of the Friedmann-Lemaıtre model it would mean general relativity theory has satisfied demanding tests on the scales of cosmology, and that we have a well-tested history of structure formation. But I would not be surprised to find that this advance leaves us with the challenge of establishing the physics of the very early universe. Acknowledgments. I thank Mike Disney, Jerry Ostriker, Michael Turner, and Neil Turok for stimulating discussions. This work was supported in part by the US National Science Foundation. References Arp, H. C. 1966, Atlas of Peculiar Galaxies (Pasadena: California Institute of Technology) Bode, P., Ostriker, J. P., & Turok, N. 2000, astro-ph/0010389 Carignan, C. & Freeman, K. C. 1988, ApJ, 332, L33 Cen, R. & Ostriker, J. P. 2000, ApJ, 583, 83 Disney, M. J. 2000, astro-ph/0009020 Hoyle, F., Burbidge, G., & Narlikar, J. V. 1993, ApJ, 410, 437 Hubble, E. 1936, The Realm of the Nebulae (New Haven: Yale University Press) Lanzetta, K. M., Bowen, D. V., Tytler, D. & Webb, J. K. 1995, ApJ, 442, 538 Peebles, P. J. E. 1989, J. Royal Astronomical Society of Canada, 83, 363 Peebles, P. J. E., Seager, S. & Hu, W. 2000, astro-ph/0004389 Popescu, C. C., Hopp, U., & Rosa, M. R. 1999, AA, 350, 414 Pustil'nik, S. A., Ugryumov, A. V., Lipovetsky, V. A., Thuan, T. X., & Guseva, N. G. 1995, ApJ, 443, 499 Sandage, A. & Bedke, J. 1988, Atlas of Galaxies (Washington: NASA Scientific and Technical Information Division), plate 11 Shull, J. M., Stocke, J. T., & Penton, S. 1996, AJ, 111, 72 Steidel, C. C., Dickinson, M., & Persson, S. E. 1994, ApJ, 427, L75 Tolman, R. C. 1934, Relativity, Thermodynamics and Cosmology (Oxford: the Clarenden Press) Tully, R. B. 1988, Nearby Galaxies Catalog (Cambridge: Cambridge University Press) Zwann, M. A., Briggs, F. H., Sprayberry, D., & Sorar, E. 1997, ApJ, 490, 173
astro-ph/0502384
1
0502
2005-02-18T21:00:50
The Magnetic Properties of an L Dwarf Derived from Simultaneous Radio, X-ray, and H-alpha Observations
[ "astro-ph" ]
We present the first simultaneous, multi-wavelength observations of an L dwarf, the L3.5 candidate brown dwarf 2MASS J00361617+1821104, conducted with the Very Large Array, the Chandra X-ray Observatory, and the Kitt Peak 4-m telescope. We detect strongly variable and periodic radio emission (P=3 hr) with a fraction of about 60% circular polarization. No X-ray emission is detected to a limit of L_X/L_{bol}<2e-5, several hundred times below the saturation level observed in early M dwarfs. Similarly, we do not detect H-alpha emission to a limit of L_{H-alpha}/L_{bol}<2e-7, the deepest for any L dwarf observed to date. The ratio of radio to X-ray luminosity is at least four orders of magnitude in excess of that observed in a wide range of active stars (including M dwarfs) providing the first direct confirmation that late-M and L dwarfs violate the radio/X-ray correlation. The radio emission is due to gyrosynchrotron radiation in a large-scale magnetic field of about 175 G, which is maintained on timescales longer than three years. The detected 3-hour period may be due to (i) the orbital motion of a companion at a separation of about five stellar radii, similar to the configuration of RS CVn systems, (ii) an equatorial rotation velocity of about 37 km/s and an anchored, long-lived magnetic field, or (iii) periodic release of magnetic stresses in the form of weak flares. In the case of orbital motion, the magnetic activity may be induced by the companion, possibly explaining the unusual pattern of activity and the long-lived signal. We conclude that fully convective stars can maintain a large-scale and stable magnetic field, but the lack of X-ray and H-alpha emission indicates that the atmospheric conditions are markedly different than in early-type stars and even M dwarfs. [abridged]
astro-ph
astro-ph
DRAFT VERSION SEPTEMBER 7, 2018 Preprint typeset using LATEX style emulateapj THE MAGNETIC PROPERTIES OF AN L DWARF DERIVED FROM SIMULTANEOUS RADIO, X-RAY, AND Hα OBSERVATIONS E. BERGER1,2,3, R. E. RUTLEDGE4, I. N. REID5, L. BILDSTEN6, J. E. GIZIS7, J. LIEBERT8, E. MARTÍN9,10, G. BASRI11, R. JAYAWARDHANA12, A. BRANDEKER12, T. A. FLEMING8, C. M. JOHNS-KRULL13, M. S. GIAMPAPA14, S. L. HAWLEY15, J. H. M. M. SCHMITT16 5 0 0 2 b e F 8 1 1 v 4 8 3 2 0 5 0 / h p - o r t s a : v i X r a Draft version September 7, 2018 ABSTRACT We present the first simultaneous, multi-wavelength observations of an L dwarf, the L3.5 candidate brown dwarf 2MASS J00361617+1821104, conducted with the Very Large Array, the Chandra X-ray Observatory, and the Kitt Peak 4-m telescope. We detect strongly variable and periodic radio emission (P = 3 hr) with a fraction of about 60% circular polarization. No X-ray emission is detected to a limit of LX /Lbol ∼< 2 × 10- 5, several hundred times below the saturation level observed in early M dwarfs. Similarly, we do not detect Hα emission to a limit of LH α/Lbol ∼< 2 × 10- 7, the deepest for any L dwarf observed to date. The ratio of radio to X-ray luminosity is at least four orders of magnitude in excess of that observed in a wide range of active stars (including M dwarfs) providing the first direct confirmation that late-M and L dwarfs violate the radio/X-ray correlation. The radio emission is due to gyrosynchrotron radiation in a large-scale magnetic field of about 175 G, which is maintained on timescales longer than three years. The detected 3-hour period may be due to (i) the orbital motion of a companion at a separation of about five stellar radii, similar to the configuration of RS CVn systems, (ii) an equatorial rotation velocity of about 37 km s- 1 and an anchored, long-lived magnetic field, or (iii) periodic release of magnetic stresses in the form of weak flares. In the case of orbital motion, the magnetic activity may be induced by the companion, possibly explaining the unusual pattern of activity and the long-lived signal. We conclude that fully convective stars can maintain a large-scale and stable magnetic field, but the lack of X-ray and Hα emission indicates that the atmospheric conditions are markedly different than in early-type stars and even M dwarfs. Similar observations are therefore invaluable for probing both the internal and external structure of low mass stars and sub-stellar objects, and for providing constraints on dynamo models. Subject headings: stars: low mass,brown dwarf -- stars: activity -- stars: radio emission -- stars: magnetic fields -- radiation mechanisms: nonthermal 1. INTRODUCTION In recent years the stellar mass function has been extended at the bottom of the main sequence and beyond with the discov- ery of numerous very low mass stars and brown dwarfs (spec- tral types late-M, L, and T; Martín et al. 1997; Delfosse et al. 1999; Martín et al. 1999; Reid et al. 1999; Luhman et al. 2000; Leggett et al. 2000; Hawley et al. 2002). Despite the significant observational and theoretical advances made since their discovery (e.g., Basri 2000; Chabrier & Baraffe 2000; Burrows et al. 2001 and references therein), many questions regarding the structure and formation of these objects remain unanswered. Not least among these is the generation, amplifi- cation and dissipation of magnetic fields and their influence on the coronae and chromospheres. In turn, these provide a win- dow onto the physics of the internal convection, the structure of the atmosphere, and its effect on the emergent radiation. By analogy with solar-type stars and M dwarfs, we expect that the magnetic activity will be evident through coronal and chromo- spheric emission (X-ray, Hα, and radio). In the M dwarfs, observations in X-rays and Hα reveal a sub- stantial fraction of active objects, peaking at ∼ 70% around 1Observatories of the Carnegie Institution of Washington, 813 Santa Barbara Street, Pasadena, CA 91101 2Princeton University Observatory, Peyton Hall, Ivy Lane, Princeton, NJ 08544 3Hubble Fellow 4Department of Physics, McGill University, Rutherford Physics Building, 3600 University Street, Montreal, QC H3A 2T8, Canada 5Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218 6Kavli Institute for Theoretical Physics, University of California at Santa Barbara, Kohn Hall, Santa Barbara, CA 93106 7Department of Physics and Astronomy, University of Delaware, Newark, DE 19716 8Department of Astronomy and Steward Observatory, University of Arizona, 933 North Cherry Avenue, Tucson, AZ 85721 9Instituto de Astrofísica de Canarias, C/ Vía Láctea s/n, E-38200 La Laguna, Tenerife, Spain 10University of Central Florida, Department of Physics, PO Box 162385, Orlando, FL 32816 11Astronomy Department, University of California, Berkeley, CA 94720 12Department of Astronomy and Astrophysics, University of Toronto, 60 St. George Street, Toronto, ON M5S 3H8, Canada 13Department of Physics and Astronomy, Rice University, 6100 Main Street, MS-61 Houston, TX 77005 14National Solar Observatory, National Optical Astronomy Observatories, Tucson, AZ 85726 15University of Washington, Department of Astronomy, Box 351580, Seattle, WA 98195 16Hamburger Sternwarte, Universitat Hamburg, Gojenbergsweg 112, 21029 Hamburg, Germany 1 hereafter, B02) revealed flares and persistent emission from three objects ranging from M8.5 to L3.5. In all cases the in- ferred magnetic fields strengths were of the order of 100 G. In addition, of the three detected objects, the two which have been observed in X-rays again violated the radio/X-ray corre- lation by orders of magnitude. Recently, Putman & Burgasser (2003) detected radio emission from four out of seven M7 to L4.5 dwarfs they observed with Australia Telescope Compact Array, of which one exhibited a strong flare. Thus, the inci- dence rate of radio emission in late-M and L dwarfs may be as high as 40%, much larger than in Hα or X-rays. Beyond the individual properties inferred from the unex- pected radio emission, two puzzling trends have emerged. First, the fraction of persistent radio emission compared to the bolo- metric luminosity appears to increase with later spectral type, the exact opposite of the trend observed in Hα and X-ray activ- ity. Second, the radio bright objects are those with high rotation velocities (vsini ∼> 20 km s- 1), while significant non-detections are associated with slow rotators (vsini ∼< 10 km s- 1). This im- plies that the rotation-activity relation holds in the radio, despite being broken in Hα and X-rays. The glaring disparity between the trends observed in the var- ious activity indicators makes it clear that a complete picture of magnetic activity in late-M and L dwarfs requires simul- taneous, multi-wavelength observations. These observations can furthermore constrain turbulent dynamo models, which de- spite significant progress are still under-developed and under- constrained. Here we present the first such observations for any L dwarf, the L3.5 object 2MASS J00361617+1821104 (here- after, 2M0036+18). In §2 we provide the details of our radio, X-ray, and Hα observations. We summarize the results in §3 and address the radio emission mechanism and derive the mag- netic field properties in §4. The violation of the radio/X-ray correlation and its implications are discussed in §5. A possible periodicity observed in the radio emission is assessed in §6 and implications for the magnetic field generation mechanism are drawn in §7. Finally, we summarize the current trends in radio, Hα and X-ray emission from M and L dwarfs in §8. 2. OBSERVATIONS We targeted the L3.5 dwarf 2M0036+18 (Reid et al. 1999) due to its relative vicinity and previous observations of radio activity. The object is located at a distance of 8.8 pc (Dahn et al. 2002), has a bolometric luminosity, Lbol ≈ 10- 3.97 L⊙ (Leggett et al. 2002; Vrba et al. 2004), a rotation velocity, vsini ≈ 15 ± 5 km s- 1, measured from high-resolution optical spectra (Schweitzer et al. 2001), an inferred temperature of about 1900 K (Vrba et al. 2004), an inferred radius of about 0.09 ± 0.01 R⊙ (Dahn et al. 2002), a mass of about 0.076 M⊙ inferred from the surface gravity, logg ≈ 5.4 (Schweitzer et al. 2001), and an inferred age of at least 1 Gyr (e.g., Burrows et al. 2001). Thus, 2M0036+18 is an object located near or below the hydrogen burning limit. 2M0036+18 has been detected in previous radio observations carried out in September and Oc- tober of 2001, with a strong ∼ 20-min flare, variable persistent emission, and a high fraction of circular polarization (B02). Surprisingly, the onset of full convection and the expected breakdown of the αΩ dynamo at spectral type M3 is not ac- companied by obvious changes in the level of Hα and X-ray emission or the rotation-activity relation. This has been at- tributed to a growing contribution from an α2 (Raedler et al. 1990) or a turbulent dynamo (Durney, De Young & Roxburgh 1993) in which the shearing needed to generate and amplify the fields comes from turbulent motion associated with the inter- nal convection. However, beyond spectral type M7 there is a precipitous drop in Hα and X-ray persistent activity, and only a handful of objects exhibit flares, ∼ 7% in late-M dwarfs and ∼ 1% in L dwarfs (e.g., Reid et al. 1999; Gizis et al. 2000; Rutledge et al. 2000; Liebert et al. 2003; West et al. 2004 and see Figure 14). Furthermore, the activity-rotation relation no longer holds in late-M and L dwarfs with many rapid rota- tors (v ∼> 20 km s- 1) exhibiting no discernible activity (Basri & Marcy 1995; Mohanty & Basri 2003). The decrease in activ- ity may be due to a quenching of the turbulent dynamo and a transition to small-scale and short-lived fields17. Moreover, the increasingly neutral atmospheres of late-M and L dwarfs may hamper the dissipation of the fields (Mohanty et al. 2002). In effect, the increased rate of collisions between charged and neu- tral particles may decouple the magnetic field from the atmo- sphere, thereby suppressing the build-up of magnetic stresses and the heating of the chromosphere and corona. The above discussion suggests that radio emission, produced by interaction of relativistic electrons with the magnetic field, should also be suppressed in late-M and L dwarfs. Observa- tionally, the empirical relation between the radio and X-ray lu- minosities of a wide range of coronally-active stars, including many M dwarfs (Guedel & Benz 1993; Benz & Guedel 1994), along with the few X-ray detections and upper limits obtained to date, suggest an expected flux level below 0.1 µJy (i.e., sev- eral hundred times lower than the detection threshold of the Very Large Array). However, radio observations of late-M and L dwarfs present a decidedly different picture. Berger et al. (2001) (hereafter, B01) detected radio flares and persistent emission from the M9.5 brown dwarf LP944-20 four orders of magnitude brighter than expected based on X-ray observations (Rutledge et al. 2000; Martín & Bouy 2002) and the radio/X-ray correlation. A subsequent survey of twelve additional objects (Berger 2002; 2 spectral type M8 (e.g., Gizis et al. 2000; West et al. 2004). Moreover, in both bands the level of activity increases with rotation velocity, up to a saturation level, at v ∼ 5 km s- 1, of LX /Lbol ≈ 10- 3 (Rosner, Golub & Vaiana 1985; Fleming et al. 1993; Pizzolato et al. 2003) and LH α/Lbol ≈ 10- 3.5 (Mohanty et al. 2002). These observations have been interpreted in the context of coronal and chromospheric heating by dissipation of magnetic fields, which in turn are powered by an internal dy- namo. In stars earlier than M3 the dynamo is thought to be pow- ered by shearing motions at the radiative-convective transition zone (the so-called αΩ dynamo). The increase in activity with decreased mass and higher rotation velocity is understood as a dependence of the dynamo on the Rossby number, Ro = P/τc, where P is the rotation period and τc is the convective turnover time (e.g., (Soderblom et al. 1993)). The saturation effect is still not fully understood. The observations presented here were carried out simultane- 17Recent work by Dobler, Stix & Brandenburg (2004) on magnetic field generation in convective rotating stars suggests that large-scale fields can be generated and maintained on relatively long timescales. However, these authors use physical conditions relevant for an M5 dwarf (and with an unusually high surface temperature), while substantial changes in Hα and X-ray activity are manifested only for objects later than M7, and we are interested in mid-L dwarfs here. ously in the radio, X-ray, and optical on September 28, 2002. Radio observations commenced at 01:49 UT and ended at 09:39 UT (28.2 ks), X-ray observations covered the range 01:10 to 07:17 UT (22 ks), and optical spectroscopic observations cov- ered the range 02:47 to 07:31 UT (17 ks). 2.1. Radio Very Large Array (VLA18) observations were conducted si- multaneously at 4.86 and 8.46 GHz in the standard continuum mode with 2 × 50 MHz contiguous bands at each frequency. We used fourteen antennas at 4.86 GHz and thirteen antennas at 8.46 GHz, with a staggered configuration in each of the three ar- ray arms. Eight-minute scans on 2M0036+18 were interleaved with 50 s scans on the phase calibrator J0042+233. The flux density scale was determined using the extragalactic source 3C 48 (J0137+331). The data were reduced and analyzed using the Astronomical Image Processing System (AIPS; Fomalont 1981). The visi- bility data were inspected for quality, and noisy points were removed. To search for flares, we constructed light curves using the following method. We removed all the bright field sources using the AIPS/IMAGR routine to CLEAN the region around each source (with the exception of 2M0036+18), and the AIPS/UVSUB routine to subtract the resulting source mod- els from the visibility data. We then plotted the real part of the complex visibilities at the position of 2M0036+18 as a func- tion of time using the AIPS/UVPLT routine. The subtraction of field sources is necessary since their sidelobes and the change in the shape of the synthesized beam during the observation re- sult in flux variations over the map, which may contaminate any real variability or generate false variability. We repeated this procedure for the background sources as a check on the in- trinsic noise properties of the maps. The resulting light curves are shown in Figures 1 and 2. We conducted follow-up observations with the VLA on 2005, Jan. 10 and 11 UT at 4.86 GHz. A total of ten hours were obtained using the same phase and flux calibrators as in the Sep. 2002 observation. We followed the same procedure outlined above to to produce light curves (Figures 8 and 9) 2.2. X-rays The observations were made with the Chandra/ACIS-S3 (backside-illuminated chip). 2M0036+18 was offset from the on-axis focal point by 4′ to mitigate pileup in the event of a flare. Data were analyzed using CIAO v3.1. We extracted data within a 4′′ circle centered on the source position, selecting counts in a low energy band (0.2-3 keV) and a total energy band (0.2-10 keV). Sources were found using celldetect. To es- timate the background we used annuli centered at the source position, excluding other point sources detected in the obser- vation. We adopt an energy conversion factor of 1 count = 4.5 × 10- 12 erg cm- 2 s- 1 (0.2-8 keV), consistent with a kT =1 keV Raymond-Smith plasma. No X-ray source is detected at the optical position of 2M0036+18. The number of counts in the low and total en- ergy bands were 3 and 5, respectively, while we expected 2 and 4 counts from the background. There is no evidence for flaring activity, or deviation from the Poisson count rate distribution 3 expected from a constant background. For example, the short- est time-span which includes 3 counts is 6.4 ks. For a detection of 5 counts, the 90% confidence limit (found from an average count rate of <9 counts per 21.5 ks) on the time-averaged flux is < 1 ×10- 15 erg cm- 2 s- 1. We observed no more than 1 count per 1 hr time period, corresponding to a 90% confidence limit on a 1 hr average peak flux (<3 counts/hr) of < 4 × 10- 15 erg cm- 2 s- 1. At the distance of 2M0036+18 the upper limit on persistent emission is LX ∼< 9.3 × 1024 erg s- 1, or log(LX /Lbol) ∼< - 4.65. 2.3. Optical 2M0036+18 was monitored spectroscopically using the Multi-Aperture Red Spectrometer (MARS) mounted on the Mayall 4-meter telescope at Kitt Peak National Observatory. A series of forty-eight 300-s exposures covering 6400 - 10,100A with a resolution of 3A pix- 1 were obtained in clear and photo- metric conditions with reasonable seeing (∼1.2′′, although the initial observations were made at airmass exceeding 2). All of the spectra were flat-fielded, extracted and wavelength cal- ibrated using standard IRAF routines. Observations of Feige 110 were used to set the flux scale. None of the spectra show evidence for Hα emission. The individual 300-s spectra have signal-to-noise of about 10 to 15 at these wavelengths, leading to an upper limit on the Hα equivalent width of ≈ 1.5A (or FH α ∼< 5 × 10- 17 erg cm- 2 s- 1). Combining all the individual spectra (Figure 3), we find no ev- idence for emission exceeding an equivalent width of 0.4A, or FH α ∼< 1 × 10- 17 erg cm- 2 s- 1. At the distance of 2M0036+18 this translates to LH α ∼< 9.3 × 1022 erg s- 1, or log(LH α/Lbol) ∼< - 6.65. Separately, on 2004 December 5 we obtained two 10-min high-resolution spectra with the Magellan Inamori Kyocera Echelle (MIKE) Spectrograph, separated by 103 min. The pur- pose of this observation was to measure a possible radial veloc- ity induced by a close-in companion (see §6). A 0.35′′ slit was used with 2 ×2 binning, providing a spectral resolution of about 27,000 in the red or about 0.11A per pixel. The spectra were flat-fielded, rectified, wavelength calibrated and extracted using a MIKE data reduction pipeline. The rms wavelength residuals were of the order of 0.005A and the signal-to-noise in the rel- evant echelle orders was approximately ten and six in the first and second spectrum, respectively. 3. BASIC PROPERTIES OF THE RADIO EMISSION While neither X-ray nor Hα emission were detected from 2M0036+18, radio emission was detected with an average flux of 259 ± 19 µJy and 134 ± 16 µJy at 4.9 and 8.5 GHz, re- spectively (Figure 1). The resulting spectral index (Fν ∝ ν α) is α = - 1.2 ± 0.3, typical of radio emission observed from M dwarfs (Gudel et al. 1993). The emission is strongly circu- larly polarized with an average of fc ≈ - 73 ± 8% at 4.9 GHz and fc ≈ - 60 ± 15% at 8.5 GHz. The negative sign indicates that the polarization direction is left-handed. We place a limit of ∼< 20% (3σ) on the fraction of linear polarization at 4.9 GHz and ∼< 35% at 8.5 GHz. In the follow-up observations we de- tected emission from 2M0036+18 at a level of 152 ±9 µJy, with a fraction of circular polarization, fc ≈ - 46 ± 7%. For comparison, two past observations of 2M0036+18 at 8.5 18The VLA is operated by the National Radio Astronomy Observatory, a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. 4 GHz (B02) revealed a similar flux level, 135 and 330 µJy, but a lower fraction of circular polarization, - 35% and - 13%. One of the two observations also uncovered a 20-min flare with a peak flux of about 720 µJy and a fraction of circular polariza- tion, fc ≈ - 62%. No such flares are observed here suggesting that their rate of occurrence, taking into account all three obser- vations, is < 0.04 hr- 1. The flare duty cycle, defined as the flare duration relative to the total observing time, is ∼< 1.5%. The brightness temperature of the radio emission, an impor- tant discriminant of the emission mechanism, is given by Tb ≈ 2 × 109Fν,mJyν - 2 GHzd2 pc(R/RJ)- 2 K, (1) where Rs/RJ ≈ 1 is the radius in units of Jupiter radii. We find that Tb ≈ 2 × 108 K if the emission is produced in the corona at a radius of R ∼ 3Rs ∼ 3RJ (Linsky & Gary 1983; Burrows, Hubbard & Lunine 1989; Leto et al. 2000) and with a covering fraction of 100%. On the other hand, if the emission is con- fined to a significantly smaller region (e.g., a coronal loop with R ∼< 0.1Rs) then Tb may exceed 1011 K. In the case of gyro- magnetic radiation (see §4.1), the inverse Compton catastrophe limit, Tb ∼< 1012 K, defines a minimum size for the emission region of about 0.08Rs ≈ 5.0 × 108 cm. 4. EMISSION MECHANISM Several radiation mechanisms can give rise to the observed radio emission (e.g., Güdel 2002): Bremsstrahlung radia- tion, gyromagnetic radiation, and coherent processes (plasma radiation or electron cyclotron maser). We can rule out bremsstrahlung radiation in this case since the expected spec- trum is flat in the optically thin regime, whereas here α ≈ - 1.2. Coherent emission processes, which have been invoked in cases of high brightness temperature (Tb > 1012 K) solar and stellar flares with strong circular polarization, are also unlikely in this case. This is primarily because the emission is expected to have a narrow bandwidth around the fundamental or second harmonic of the plasma frequency, νp ≈ 9000n1/2 e Hz, or the cy- clotron frequency, νc ≈ 2.8 × 106B Hz, whereas here emission is detected at both 4.9 and 8.5 GHz. In addition, the typical emission timescale of coherent radiation flares is much shorter than the 7.8-hour duration of the radio emission detected in this case. 4.1. Gyromagnetic Radiation Gyromagnetic radiation can arise from a thermal or non- thermal distribution of electrons, with the latter typically as- - p, above some cut- sumed to follow a power law, n(γ) ∝ γ off, γm. The emission properties further depend on the har- monic number, s, or equivalently the typical energy of the elec- trons γ = s1/3. Cyclotron emission is produced when s ∼< 10 (non-relativistic electrons, typically thermal distribution), gy- rosynchrotron when s ≈ 10 - 100 (mildly relativistic electrons), and synchrotron when s ∼> 100 (relativistic electrons). The fre- quency at which the bulk of the radiation is emitted is given by νm ≈ 2.8 × 106sB Hz, where B is the magnetic field strength. Cyclotron radiation is unlikely in this case since the emis- sion is expected to be concentrated in narrow bands at the har- monic frequencies, as opposed to the detected signal at 4.9 and 8.5 GHz. Similarly, gyrosynchrotron emission from a thermal plasma is unlikely since the expected optically thin spectral in- dex, α ≈ - 8, is significantly steeper than the observed value. Finally, synchrotron emission is unlikely because the expected degree of circular polarization is strongly suppressed compared to the value of about 70% measured here. Therefore, the most likely emission mechanism in the case of 2M0036+18 is gy- rosynchrotron radiation from a non-thermal distribution of elec- trons. The emission properties in the gyrosynchrotron case depend both on the fundamental source parameters -- magnetic field B, electron density ne, and line-of-sight source scale L -- and the angle between the line of sight and the magnetic field, θ (Dulk & Marsh 1982). Since the latter cannot be inferred from the present data we take θ = π/3 as a representative value. The spectral index in the optically thin case, α ≈ 1.2 - 0.9p, indi- cates that p ≈ 2.7. This value is similar to those found in sev- eral early M dwarfs (Gudel et al. 1993). Given the value of p, the fraction of circular polarization at 4.9 GHz is given by fc ≈ 0.025B0.53 ≈ 0.6, from which we infer B ≈ 400 G. For the full range of possible θ values the magnetic field strength ranges from about 80 to 1.9 × 103 G. These values indicate that the harmonic number is s < 22, with the gyrosynchrotron limit, s > 10, corresponding to B < 175 G, or θ < π/4. Below we use the latter value for the magnetic field strength. The frequency at which the gyrosynchrotron spectrum peaks is given by νm ≈ 6.1 × 105(neL)0.24 < 4.86 × 109 Hz, while the flux in the optically thin regime is Fν ≈ 1.1 × 10- 34neLR2 ≈ 205 µJy, using the average flux at 4.9 GHz. Solving for the un- known column density and radius of the emission region we find neL < 1.8 ×1016 cm- 2 and R ≈ 1 ×1010 cm, or about 1.4Rs. The resulting brightness temperature is Tb ≈ 7 × 108, consis- tent with the interpretation of the radio emission as gyrosyn- chrotron radiation. Assuming that L = R, the electron density is ne ≈ 1.6 × 106 cm- 3. Several conclusions can be drawn from this discussion. First, the magnetic field strength is larger than the average solar field or the magnetic field of Jupiter, but is somewhat lower than the ∼ 1 kG fields inferred in some active early M dwarfs (Smith et al. 1975; Saar & Linsky 1985; Haisch, Strong & Rodono 1991; Johns-Krull & Valenti 1996; Stepanov et al. 2001). We note however, that the surface magnetic field of 2M0036+18 may be an order of magnitude larger than that inferred for the emission region if L ≈ R ≈ 1.4Rs, placing it in the same range of early M dwarfs. Second, the radio emission, and in particular the high fraction of circular polarization, requires a large-scale, long-lived ordered magnetic field with a large covering frac- tion. This is again similar to the conditions inferred for early M dwarfs (Saar & Linsky 1985; Johns-Krull & Valenti 1996). Fi- nally, the density of emitting electrons is similar to that inferred in the emission zones (coronae) of early M dwarfs (Stepanov et al. 2001; Osten et al. 2004). Thus, the physical conditions in the L3.5 2M0036+18 resemble those found in active early M dwarf, but the Hα and X-ray emission are strongly suppressed. We finally note that the plasma-beta, β = 2µ0P/B2, of the corona, using n ∼ 1.6 ×106 cm- 3, a typical temperature T ∼ 2 × 106 K (Giampapa et al. 1996), and B ∼ 175 G, is β ∼ 3 × 10- 8. The typical photospheric pressure of an L dwarf is P ∼ 3 × 105 dyne cm- 2 (Burrows et al. 2001), resulting in β ∼ 2. These con- ditions are again similar to those found in M dwarfs (Giampapa et al. 1996) and the Sun. Thus, even small perturbations in the coronal field of 2M0036+18 are capable of supporting the pressure gradients produced by the radio-emitting material. 5. VIOLATION OF THE RADIO/X-RAY CORRELATION The simultaneous radio and X-ray observations allow us for the first time to directly investigate the radio/X-ray correla- tion in an L dwarf. Similar observations of coronally active stars (including the Sun) reveal a tight correlation between LR and LX (Figure 4; Guedel & Benz 1993; Benz & Guedel 1994). This correlation holds for both flaring and persistent emission, and for single stars and binary systems. For M dwarfs the correlation holds to spectral type M7 (Gudel et al. 1993). The persistent emission follows LR ≈ 3 × 10- 16LX Hz- 1 (Guedel & Benz 1993), and extends over six orders of magni- tude in LR. The relation for radio and X-ray flares is non-linear, LR ∝ 1.3 × 10- 26L0.73 X Hz- 1 (Benz & Guedel 1994), and holds over eight orders of magnitude in LR. For 2M0036+18 we find LR ≈ 2.5 × 1013 erg s- 1 Hz- 1 and LX < 9.7 × 1024 erg s- 1. Thus, LR/LX ∼> 2.6 × 10- 12 Hz- 1 is at least 8600 times higher than predicted by the radio/X-ray corre- lation (Figure 4). A similar result, from non-simultaneous data, was obtained for the M9.5 brown dwarf LP944-20 (B01) with a ratio that exceeded the expected value by a factor of 104 (flares) and > 8500 (persistent emission). Similarly, for the M9.5 dwarf BRI 0021-0214 the radio luminosity exceeded the X-ray upper limit by at least a factor of 103 (B02). Thus, the radio and X-ray luminosities appear to de-correlate over a narrow range in spec- tral type (between about M7 and M9) and remain uncorrelated at least to mid-L (Figure 4). Incidentally, this is the same range over which both the fraction of Hα and X-ray active sources and the strength of the activity drop precipitously. The violation of the radio/X-ray correlation is surprising given that the magnetic and emission region properties derived for 2M0036+18 are not dissimilar from those found on early M dwarfs (§4.1). This suggests that either the origin of the magnetic field is different than in M dwarfs or the atmospheric conditions are not conducive for the support of a chromosphere and/or a corona. The radio/X-ray correlation has been interpreted in the con- text of magnetic reconnection. In this scenario, the flares are produced when coronal magnetic loops reconnect, release en- ergy, and create a current sheet along which ambient electrons are accelerated. The accelerated electrons drive an outflow of hot plasma into the corona as they interact with and heat the underlying chromospheric material. The interaction of the out- flowing plasma with the electrons produces X-ray emission via the bremsstrahlung process (Neupert 1968; Hawley et al. 1995; Guedel et al. 1996). This so-called Neupert effect points to a causal connection between particle acceleration, which is the source of radio emission, and plasma heating, which results in X-ray emission. Thus, the X-ray thermal energy should sim- ply be related by a constant of proportionality to the integrated radio flux. Guedel & Benz (1993) discuss this correlation in terms of the relation LR/LX ≈ 10- 22B2.5(a/b)τ0(1 + α) Hz- 1, where a is the fractional efficiency of accelerating electrons, b is the fraction of coronal energy radiated in X-rays, τ0 is the average lifetime of the low-energy electrons, and α is the power law relating the electron lifetime to its energy. The observed tight correlation suggests that the combination B2.5(a/b)τ0 is nearly unchanged for a wide range of stars. Since in the case of 2M0036+18 and the other radio active late-M and L dwarfs LR/LX is higher by several orders of magnitude while the inferred value of B is 5 not unusual compared to M dwarfs, this indicates that either the fraction of energy emitted in X-rays, b, is suppressed by a factor of ∼ 104, or the typical lifetime of the electrons, τ0, is longer by a similar factor. In particular, for a/b of order unity and B ≈ 175 G, we find a mean electron lifetime τ ∼ 1 d. In this case, the X-ray emission is suppressed because the electrons do not lose their energy to coronal heating on a sufficiently rapid timescale. On the other hand, the violation of the radio/X-ray correla- tion may be due to an inefficient production of X-rays, i.e., a low value of b. In the case of L dwarfs this may be explained by the increasingly neutral atmospheres and the reduction in the density of ions available for X-ray emission (Mohanty et al. 2002). However, it is not clear if this can account for the late-M dwarfs LP944-20 and BRI 0021-0214, or for the sud- den transition between spectral types M7 and M9. More likely, the reduction in X-ray luminosity, or the over-production of ra- dio emission, may be related to the magnetic structure across the atmosphere, chromosphere and corona or efficient trapping of the electrons (i.e., long lifetime). We note that recent radio and X-ray observations of the dM4.5e star EV Lac also reveal a breakdown of the Neupert effect, which Osten et al. (2004) interpret in the same context discussed here, namely efficient trapping or a low value of b. 6. VARIABILITY AND PERIODICITY To this point we have discussed the properties of the inte- grated radio emission during the entire 7.8 hour simultaneous observation and the follow-up observations. However, from Figures 1, 2, 8, and 9 it is clear that both the total intensity and circularly polarized emission at 4.9 GHz are strongly vari- able. For the total intensity light curve the reduced χ2 assum- ing a constant flux is ≈ 2.5 - 7.5 for a time resolution ranging from 5 min to 1 hour; for the circular polarization light curve r ≈ 1.8 - 2.8 over the same range of time resolutions. It is χ2 not clear if the variability of the weaker emission at 8.5 GHz is r ≈ 1 - 1.7 depending on the time resolution. significant, with χ2 We are confident that this variability is not an observational artifact for two reasons. First, the light curves of two field sources, constructed in the same manner as that of 2M0036+18 (see §2.1), do not exhibit significant variability, with χ2 r ≈ 1.2 and ≈ 1.8 for a time resolution of 10 min. In fact, the re- duced χ2 values for the field sources are over-estimated com- pared to those for 2M0036+18 since, unlike 2M0036+18, these sources are spatially extended and positioned away from the phase center. As a result, changes in the synthesized beam dur- ing the observation affect their light curves more significantly than that of 2M0036+18. Second, the variability of 2M0036+18 is significant in circular polarization where no field sources, which can introduce false variability, produce detectable emis- sion. It is important to note that in the two past observations of 2M0036+18, lasting 150 and 180 minutes, the persistent emis- sion was also variable (B02). In addition to the variability, it appears from Figures 2, 8, and 9 that the radio emission at 4.9 GHz is periodic. The auto- correlation function (ACF) of both the total intensity and circu- lar polarization at a wide range of time resolutions shows a clear peak at 3 hours (Figures 5 and 10). The ACFs of the two field sources described above are flat, as expected for non-periodic sources. We assess the statistical significance of the 3-hour periodic- 6 ity using two other methods which provide a measure of the power spectrum and are therefore more easily calibrated: The Lomb-Scargle (LS) periodogram (e.g., Press et al. 1992), and a one-dimensional CLEAN algorithm (Roberts, Lehar & Dreher 1987). The latter method has the added advantage that artifacts and spectral leakage which arise from the sampling function and the finite length of the observation are removed. We assess the significance of the results using the Monte Carlo method. Specifically, we construct both randomized ver- sions of the 4.9 GHz light curves, and simulated light curves with a (χ2) variability similar to that of 2M0036+18. Depend- ing on the time resolution we use 50,000 to 100,000 light curves in each set and perform the same procedures as on the real data. The significance of peaks in the LS and CLEAN power spectra of the real light curve are determined as the frac- tion of simulated light curves which produce stronger peaks. We find that both the randomized and simulated sets provide a similar measure of the significance, as do the LS and CLEAN procedures. A representative LS periodogram is shown in Fig- ure 6. The significance of peaks in the LS and CLEAN power spec- tra does depend on the choice of time resolution. This is ex- pected since as the period is sampled more coarsely the strength of the signal is expected to decrease. We use time resolu- tions ranging19 from 2 to 30 min (Figure 7), and find that the strongest peak always occurs at a frequency of 0.33 hr- 1, or a period of 3 hours. For the simulated light curves the peaks that exceed the value for 2M0036+18 occur over the entire fre- quency range. We also find that the significance of the peri- odicity in the circular polarization light curve is consistently higher than for the total intensity light curve. This is expected because the contamination from field sources is non-existent in circular polarization. For the range of time resolutions investi- gated we find that the significance of the 3-hour period peaks at ≈ 99.999% for δt = 2 min. This result is confirmed by the follow-up observations, and a phased light curve of the data from 2005, Jan. 10 and 11 UT, folded with a period of 184 min is shown in Figure 11. Thus, the radio emission from 2M0036+18 is periodic with P = 3 hours. This period is maintained on timescales of at least two years. If related to the rotation of 2M0036+18 then using Rs ≈ 6.3 × 109 cm, the inferred period indicates a surface equa- torial rotation velocity v ≈ 37 km s- 1. This value is nearly 2.5 times higher as that inferred from high-resolution optical spec- tra, vsini ≈ 15 ± 5 km s- 1 (Schweitzer et al. 2001), suggesting an inclination angle, i ≈ 24 ± 8◦. Alternatively, if the period is related to the orbital motion of a companion which induces magnetic activity (see below) then the semi-major axis is about 3.1 × 1010 cm or about five times the stellar radius; this is simi- lar to the case of the highly active RS CVn systems (e.g., Mutel & Lestrade 1985). Finally, the periodic rise in flux may be a series of weak flares, rather than variable persistent emission, in which case the flare generation process is periodic. 7. THE ORIGIN OF THE MAGNETIC ACTIVITY IN 2M0036+18 The results presented in the previous sections directly con- firm for the first time that the relative activity patterns in late-M and L dwarfs differ from those in early-type stars and early M dwarfs. It is therefore crucial to address the origin of the mag- netic fields in the well-studied 2M0036+18 as an example of a larger trend that is emerging in sub-stellar objects. While some dynamo models for convective stars predict a reduced strength, physical scale, and lifetime of the magnetic field (e.g., Durney, De Young & Roxburgh 1993), the extent of this reduction is still not known. Clearly, the stability of the magnetic field of 2M0036+18 over a period of three years indicates a long-lived process. Similarly, the physical scale of the field, R ∼ Rs, is sig- nificantly larger than the physical scale of the convection. Fi- nally, the inferred field strength and electron densities are sim- ilar to those inferred in early M dwarfs. It is therefore possible that the current turbulent dynamo theories are missing some key ingredients, or that the magnetic fields are generated and ampli- fied in another process, possibly by interaction with a close-in companion. 7.1. The Influence of a Close-in Companion The enhancement of magnetic fields by a close companion may play a role in the RS CVn class of active short-period bi- naries, and a comparison is therefore particularly illustrative. These systems are known to produce radio gyrosynchrotron emission from magnetic fields of about 200 G, with a typical luminosity of Lν,R ≈ 2 × 1016 erg s- 1 Hz- 1, a factor of 103 higher than that from 2M0036+18. Using the scaling for gy- rosynchrotron emission, Lν,R ∝ B- 3/4R2 s , and taking into ac- count the similar magnetic field strengths and covering frac- tions, as well as the ratio of radii (0.09 R⊙ vs. ∼ 2 R⊙), we find Lν,2M/Lν,RS CVn ≈ 2 × 10- 3, in close agreement to the mea- sured ratio. In addition, the ratio Lν,R/Lbol for RS CVn sys- tems is known to be correlated with the orbital period of the system, such that tighter binaries are more radio active (Mu- tel & Lestrade 1985; Drake, Simon & Linsky 1989). Extrap- olating this relation to the value Lν,R/Lbol ≈ 6 × 10- 17 Hz- 1 for 2M0036+18 we predict an orbital period of about 1 hour. This is not so dissimilar from the period of 3 hours inferred here. Thus, it is possible that 2M0036+18 represents a low- mass scaled version of the RS CVn binaries and that the same type of interaction that enhances the magnetic fields in RS CVn systems may take place here. Another illustrative example is the recent detection of en- hanced chromospheric Ca II H & K emission from the star HD 179949 orbited by a 0.84 MJ planet with an orbital period of about 3.09 days, or a semi-major axis of 6.7 × 1011 cm (Shkol- nik, Walker & Bohlender 2003). The emission appears to be periodic and in phase with the orbit of the planet. The conclu- sion drawn by Shkolnik, Walker & Bohlender (2003) is that the planet induces magnetic activity on the stellar surface which gives rise to chromospheric emission. These diverse examples illustrate that interaction with a com- panion may enhance activity. In this context both tidal and mag- netic effects can play a role (Cuntz, Saar & Musielak 2000). In the case of tidal interaction, if the orbital and rotation ve- locities are not the same, the generation of time-varying tidal bulges may give rise to increased turbulent motions or an en- hanced α-effect. The synchronization timescale (e.g., Goldre- ich & Soter 1966) for a companion orbiting 2M0036+18 is s /GMs)(Ms/Mp)2(a/Rs)6 ≈ 1.6Ms,J Gyr; a long, tsync ≈ Qsω(R3 subscript p designates the primary, a subscript s designates the secondary, ω is the angular velocity, and a is the orbital separa- 19We do not investigate time resolutions smaller than 2 min since the computation time becomes prohibitively long. tion. Thus, it is likely that the putative companion is not tidally locked and tidal interaction may be an effective mechanism. The strength of the tidal effect depends sensitively on both the binary separation and the mass ratio, ∆gp/gp ∝ (Ms/Mp)[Rp/(a - Rp)]3. This suggests that the tidal effect may be particularly effective for late-M and L dwarfs where the pri- mary mass is low and the mass ratio may thus be of order unity. For example, for the HD 179949 system the distortion is ∆gp/gp ≈ 1.3 × 10- 6, while in the case of 2M0036+18 as- suming a Jupiter-mass companion and using a ≈ 5Rp we find ∆gp/gp ≈ 3 × 10- 4. m,pd2 Direct magnetic interaction is also possible, with the mag- netic energy of the system proportional to the product of the primary and secondary field strengths (Cuntz, Saar & Musielak 2000). In this case the energy is EB ∝ BpBs/d2 m,s, where dm,s is the magnetospheric radius of the secondary and dm,p ≡ a - dm,s is the distance from the primary to that radius. The en- ergy flux is related to EB via F ∝ ǫEBvc, where ǫ is an efficiency factor and vc is the combined velocity of the stellar magnetic motions (or turbulence) and the relative velocity of the orbit (at the stellar surface) compared to the rotation of the star. Thus, as in the case of tidal interaction, the possible larger mass ratio and small separation in the case of 2M0036+18 suggests that this mechanism may be more efficient than in extra-solar plan- ets orbiting solar-type stars. A final possibility is that a brown dwarf companion to 2M0036+18 may undergo sustained Roche lobe overflow which may result in enhanced activity (Burgasser et al. 2000, 2002). This scenario has been proposed for Hα active dwarfs, although no direct evidence was found to support this hy- pothesis. Here we may for the first time have the required evidence. For sustained overflow to occur, the logarithmic change in radius of the secondary, dlnRL/dlnMs < - 1/3, re- quires a mass ratio q ≡ Ms/Mp < 0.63. The period is given by P2 = 4π2a3/GMp(1 + q) while the radius of the secondary is equal to the Roche lobe radius, Rs = RL = 0.49aq2/3/[0.6q2/3 + ln(1 + q1/3)]. The radius of the secondary is also given by Rs = (4/3)π(qMp)- 1/3/[1 + 1.8(qMp)- 1/2]4/3 (Zapolsky & Salpeter 1969; Stevenson 1991), where the constant of 4/3 is appropri- ate for an age of 1 - 5 Gyr (Burgasser et al. 2002). Using P = 3 hr and Mp = 0.076 M⊙ we find q ≈ 0.14, consistent with sustained mass transfer, a secondary mass Ms ≈ 13 MJ, a sec- ondary radius Rs ≈ 1.1 RJ, and an orbital separation a ≈ 5.2Rp. the hypothesis of a close-in companion? As in the case of planetary com- panions around solar-type stars, radial velocity, astromet- ric shift, and photometric techniques may provide an an- swer. The expected radial velocity amplitude (assuming zero eccentricity) in the case of 2M0036+18 is A ≈ 2.3 × 103(P/3 hr)- 1/3(Mssinio/MJ)(Mp/0.076M⊙)2/3 m s- 1 for a Jupiter-mass companion. Using a cross-correlation of the four strongest and narrowest absorption features of 2M0036+18 (Figure 12) we obtain a limit of A < 4 km s- 1 from the pair of MIKE spectra separated by about half a period (Figure 13). This indicates that Mssinio < 1.7 MJ if the limit is on the max- imum amplitude. Naturally, if the two epochs were obtained when the putative companion was located ∼ 90◦ away from the line of sight to 2M0036+18, the lack of a radial velocity signa- ture does not provide a useful limit. For an orbital inclination cosio < (Rp + Rs)/a, or io ∼> 68◦, Is there a way to directly test 7 we expect to observe an eclipse as the companion occults 2M0036+18 (io = 0◦ corresponds to a face-on orbit). For io ∼> 73◦ more than half the surface of 2M0036+18 would be occulted resulting in large photometric variability. If all incli- nations are equally likely, the probability of eclipse is about 24%. Gelino et al. (2002) find no significant variability for 2M0036+18 in their I-band study of photometric variability in L dwarfs. Observations were obtained with a time separation as short as 1 - 2 hours, while the maximum time baseline in- vestigated was 53 days. The lack of variability indicates that io < 68◦ if 2M0036+18 is in fact orbited by a close companion. Finally, a companion will induce an astrometric shift given by θ ≈ 0.1MsM2/3 (P/3 hr)2/3 mas for a 2 : 1 mass ratio. This p is just within the reach of current Very Long Baseline Interfer- ometry (VLBI) observations, but lower mass ratios will result in an undetectable shift. In principle, an astrometric shift down to a mass ratio of 50 : 1, or Ms ≈ 1.7 MJ, may be detected with the 4 µas precision of the Space Interferometry Mission (SIM). 7.2. Magnetic Structure and Rotation The 3-hour period may alternatively be due to the rotation of 2M0036+18. Using the inferred radius of 0.09 R⊙, the equa- torial velocity is 37 km s- 1, leading to an inclination angle of 24 ± 8◦, or nearly pole-on. The azimuthal orientation, φ, of the axis of rotation relative to the line of sight is not known. How- ever, given the low inclination and the large covering fraction, the radio-emitting region has to be located at a low latitude for most φ values. This is because a location near the pole would make the bulk of the region visible throughout the rotation pe- riod, resulting in a nearly constant flux level. As can be seen from Figure 2 the flux drops to nearly zero between the peaks indicating the emission region is fully occulted for about half the rotation period. We note also that the low inclination may explain the constant sign of the circular polarization since for most values of φ only one of the hemispheres is visible. If the 3-hour period is related to the rotation of 2M0036+18 then we can estimate the Rossby number, Ro = P/τc, which is relevant for dynamo models. The convective turnover time for 2M0036+18 is τc = (MR2/L)1/3 ≈ 0.8 yr indicating that Ro ≈ 4.4 × 10- 4. At such low values early M dwarfs exhibit sat- urated X-ray emission, LX /Lbol ∼ 10- 3 (Pizzolato et al. 2003). Clearly, the conditions in L dwarfs are sufficiently different that a low Rossby number does not result in X-ray emission. We also note that the large value of τc may explain the stability of the radio emission (flux and circular polarization) over a period of about three years. 7.3. Periodic Flaring Finally, it is possible that the variable persistent emission is in fact a series of periodic flares with a rate of occurrence of about 0.33 hr- 1. The frequency of strong flares, similar to the one detected by B01 with a flux of about 720 µJy, is ∼< 0.04 hr- 1. This pattern is similar to the case of the M9 dwarf LHS 2065 for which strong Hα flares have an occurrence rate of ∼< 0.03 hr- 1 while weak flares occur at a rate as high as about 0.5 hr- 1 (Martín & Ardila 2001). If this is the case here, then the timescale to build up magnetic stresses is about 3 hours, while the lifetime of the electrons, corresponding to the width of the flares, is about 1 hour. In this scenario it is possible that the periodic weak flares 8 are inefficient at heating the corona and chromosphere, thus ex- plaining the lack of accompanying X-ray and Hα emission. The more rare strong flares, on the other hand, with an impulsive en- ergy release a few times larger may be accompanied by efficient heating. However, we note that the time-integrated energy re- lease in the putative flares observed here, E ∼ νLνt ∼ 4 × 1026 erg, is not so different from the integrated energy release in- ferred for the stronger flare (B02). 8. EMERGING ACTIVITY PATTERNS OF LATE-M AND L DWARFS The growing sample of late-M and L dwarfs observed in the radio and X-rays allows to investigate the patterns of activity as a function of relevant physical parameters. Hα observations of late-M and L dwarfs indicate a significant drop in the level of emission compared to the bolometric luminosity (Figure 14). Objects earlier than about M6 tend to have a ratio LH α/Lbol in the range of 10- 4 to 10- 3. However, by spectral type L0 this ratio is typically lower than 10- 5. The few late-M and L dwarfs which exhibit flares (∼ 1%; Liebert et al. 2003) appear to have levels of emission similar to those of early M dwarfs, with only a single source exceeding that level to date. Thus, Hα emis- sion drops significantly with a decreased surface temperature (later spectral type), and flares exceeding the saturation value are extremely rare. Similarly, the ratio LX /Lbol drops from a saturation value of about 10- 2.5 in mid-M dwarfs to less than 10- 4 for the few late- M dwarfs observed to date (e.g., Fleming et al. 1993; Rut- ledge et al. 2000; Martín & Bouy 2002). With the exception of a recently-discovered M9 dwarf for which a ratio of about 0.1 was measured during a flare (Hambaryan et al. 2004), the flares observed in other late-M dwarfs do not exceed the saturation level observed in the mid-M dwarfs (Stelzer 2004). Thus, in both Hα and X-rays the saturation effect appears to be real, with flares replacing persistent emission in late-M and L dwarfs. 2M0036+18 appears to follow the general trend with upper limits on the X-ray and Hα emission that are about three orders of magnitude below the respective saturation values. On the other hand, the radio activity in late-M and L dwarfs, and in particular 2M0036+18, exhibits a completely different trend. The observed ratio LR/Lbol is about 10- 8 for persistent emission from early and mid-M dwarfs. However, the level of radio emission from late-M and L dwarfs exceeds this value by about an order of magnitude, while flares are brighter by at least two orders of magnitude. 2M0036+18 in particular has a ratio LR/Lbol of about 10- 6.2. Thus, while a significantly smaller number of objects have been observed in the radio compared to Hα, every single detection is brighter than the value observed in mid-M dwarfs. Moreover, the fraction of detected objects is much higher, ∼ 40%. This suggests that a saturation effect does not exists in the radio band. Moreover, the radio luminos- ity does not seem to decrease with surface temperature. Another important trend, proposed by B02, is that rapid rota- tors exhibit stronger radio activity. This is shown in Figure 15. In Hα there is a clear rotation-activity relation in dwarfs earlier than M7, but this relation clearly breaks down in late-M and L dwarfs. A similar relation exists in X-rays, but it again breaks down in late-M and L dwarfs. In the radio band, on the other hand, the rotation-activity relation appears to extend to late-M and L dwarfs. The only objects for which the ratio of radio to bolometric luminosity is lower than the detected objects are those with vsini ∼< 10 km s- 1. Thus, it appears that the efficiency of magnetic field genera- tion and dissipation in fact does not decrease in late-M and L dwarfs, but the efficiency of chromospheric and coronal emis- sion does. Similarly, the rotation activity relation appears to still hold, but the activity is manifested in radio emission instead of Hα or X-rays. The detection of rare Hα and X-ray flares does suggest that chromospheric and coronal heating may still occur, but this may require large reconnection events, possibly accom- panied by large radio flares. 9. CONCLUSIONS The simultaneous, multi-wavelength observations presented in this paper provide unparalleled insight into the magnetic field properties of an L dwarf. Most importantly, these observations directly confirm that radio activity is more prevalent in late-M and L dwarfs compared to Hα and X-ray activity. The detec- tion rate in the radio may be as high as 40%, while in Hα it is at most a few percent. Only a few objects later than M7 have been observed in the X-rays, but persistent emission is clearly rare. A detailed analysis of the radio emission along with the lack of detectable X-ray and Hα emission give rise to the following inferences: 1. The radio emission from 2M0036+18, as compared to is the strongest observed the bolometric luminosity, in any dwarf star to date. Moreover, the emission from 2M0036+18 violates the radio/X-ray correlation by at least four orders of magnitude suggesting a dras- tic change in the coronal and chromospheric conditions compared to those observed in active stars down to spec- tral class M7. This is the first direct confirmation of this violation which has been observed in two additional late- M dwarfs (B01; B02). 2. The radio emission is due to gyrosynchrotron radiation in a magnetic field of about 175 G. The size of the emis- sion region is about 1010 cm (∼ 1.4Rs), indicating a large covering fraction at the surface. These characteristics are similar to those inferred in early M dwarfs. 3. The constant fraction and sense of the circular polariza- tion in observations conducted over a period of more than three years, implies that an ordered magnetic field is maintained stably on long timescales, despite the oc- casional production of strong flares (B02); the incidence rate for the latter is < 0.04 hr- 1. 4. The radio emission is periodic (∼> 99.999% confidence level) with a period of 3 hours. If related to the rota- tion period this indicates a rotation velocity of about 37 km s- 1, while if related to an orbital period it indicates a semi-major axis of about five times the stellar radius. Alternatively, the period may be due to periodic release of the magnetic energy in the form of weak radio flares. Clearly, the radio emission from 2M0036+18 requires a large-scale, long-lived, and relatively strong magnetic field. The lack of accompanying Hα and X-ray emission suggests that the general decline in these activity indicators in late-M and L dwarfs is not a reliable tracer of decreased magnetic activity. In particular, if the magnetic field is essentially decoupled from gas in the cool atmospheres of late-M and L dwarfs so that magnetic stresses are suppressed, this may quench the chro- mospheric and coronal emission. However, at larger distances the rate of collisions between charged and neutral particles may drop and the field could then couple effectively to the gas (Mo- hanty et al. 2002). This is the region where radio emission is expected to be generated most effectively, explaining the strong radio signal compared to the reduced Hα and X-ray emission. However, even this scenario cannot easily account for the severe violation of the radio/X-ray correlation, or for the detection of Hα and X-ray flares from some L and T dwarfs. While the origin of the observed periodicity remains un- known, all three scenarios provide interesting constraints for turbulent dynamo models. In the case of periodicity due to a close-in companion, it is possible that the magnetic field of 2M0036+18 is amplified by interaction with the compan- ion. This may alleviate the need to generate strong, long-lived fields from convection alone. If the incidence of activity in the radio band is truly 40%, this may also require an unusu- ally high fraction of companions, especially in comparison to current estimates of the binary fraction in late-M and early L dwarfs of ∼ 15% (Close et al. 2003; Gizis et al. 2003). How- ever, this fraction is estimated primarily from direct detection with AO systems and the Hubble Space Telescope which are only sensitive to companions at separations of ∼> 1 AU, sev- eral hundred times larger than the possible orbit in the case of 2M0036+18. Direct evidence for close-in companions is re- quired before strong claims can be made, but radio active late- M and L dwarfs may provide a signpost for binary systems. On the other hand, in the case of periodicity due to rotation, it is clear that any dynamo model has to explain the presence 9 of magnetic structures similar to those of M dwarfs without the benefit of a radiative-convective transition zone and without hy- drogen burning as a source of heat. Obviously, if the observed radio emission is in fact due to a series of episodic flares, this places stringent constraints on the timescale of magnetic field amplification and dissipation (∼ 1 hour) over a large fraction of the stellar surface. The multi-wavelength approach taken in this paper is clearly the first step in addressing the details of magnetic field gener- ation at the bottom of the main sequence and beyond. Contin- ued studies, particularly in the radio and X-ray poorly-sampled range M9 to L5 (Figure 14), are essential for understanding in- dividual objects, as well as the effects of various physical prop- erties (e.g., binarity, rotation, temperature, age). These obser- vations will for the first time provide robust constraints for dy- namo models and for the structure of late-M and L dwarfs all the way from the interior to the corona. We thank first and foremost Barry Clark (NRAO), Scott Wolk (CXC, HEAD), and Meghan McGarry (CXC, HEAD) for their efforts in scheduling the simultaneous VLA and Chandra observations. Support for this work was provided by NASA through Chandra Award Number GO2-3012B issued by the Chandra X-Ray Observatory center, which is operated by SAO for and on behalf of NASA under contract NAS8-39073. Fur- ther support was provided by National Science Foundation grant NSF PHY 99-07949. EB is supported by NASA through Hubble Fellowship grant HST-01171.01 awarded by the Space Telescope Science Institute, which is operated by AURA, Inc., for NASA under constract NAS 5-26555. APPENDIX X-RAY OBSERVATIONS OF THE L5 DWARF 2MASS J15074769-1627386 In addition to the X-ray observations of 2M0036+18 outlined in this paper, we also observed the L5 dwarf 2MASS J15074769- 1627386 (hereafter, 2M1507-16; Reid et al. 2000; Kirkpatrick et al. 2000) as part of our Chandra AO3 program. This object is located at a distance of 7.4 pc (Reid et al. 2000, 2001; Dahn et al. 2002), has a surface temperature of about 1630 K, a bolometric luminosity of 10- 4.27 L⊙, an inferred radius of 0.09 R⊙ (Vrba et al. 2004), and a rotation velocity vsini ≈ 27 km s- 1 (Bailer-Jones 2004). The upper limit on Hα emission is LH α/Lbol < 10- 5.76 (Reid et al. 2000) and no Li is detected to a limit of 0.1A. Finally, B02 find an upper limit on the radio luminosity at 8.5 GHz of 3.8 × 1012 erg s- 1 Hz- 1 from a 2.2-hr observation, corresponding to LR/Lbol ∼< 10- 6.8. The details of the X-ray observation and data analysis are given in §2.2. No X-ray source is detected at the position of 2M1507-16. The number of counts in the low and total energy bands were 2 and 4 respectively, while we expect 2 and 5 counts from background alone. We find no evidence for variability: In the total energy band, the shortest time-span which includes 3 counts is 2330 s, which we find from a Monte Carlo simulation, will occur in 22% of 27.7 ks light curves containing four counts. The 90% confidence limit on the time-averaged flux (<7 counts per 27.7 ks) is < 1 × 10- 15 erg cm- 2 s- 1. Taking the 3 counts within 2330 s, we place a 90% confidence upper limit on a 1-hr peak flaring flux of < 1 × 10- 14 erg cm- 2 s- 1. At the distance of 2M1507-16, the upper limit on persistent emission is LX ∼< 6.6 × 1024 erg s- 1, or log(LX /Lbol) ∼< - 4.5. The upper limits on X-ray, radio, and Hα activity for 2M1507-16 are shown in Figures 14 and 15. Clearly this object follows the general trend of decreased Hα and X-ray activity in L dwarfs. The radio activity is at least a factor of four lower than that of 2M0036+18, but the upper limit is several times brighter than other late-M and L dwarfs with a similar rotation velocity. Bailer-Jones, C. A. L. 2004, A&A, 419, 703. Basri, G. 2000, ARA&A, 38, 485. Basri, G. and Marcy, G. W. 1995, AJ, 109, 762. References 10 Benz, A. O. and Guedel, M. 1994, A&A, 285, 621. Berger, E. 2002, ApJ, 572, 503. Berger, E. et al. 2001, Nature, 410, 338. Burgasser, A. J., Kirkpatrick, J. D., Reid, I. N., Liebert, J., Gizis, J. E., and Brown, M. E. 2000, AJ, 120, 473. Burgasser, A. J., Liebert, J., Kirkpatrick, J. D., and Gizis, J. E. 2002, AJ, 123, 2744. Burrows, A., Hubbard, W. B., and Lunine, J. I. 1989, ApJ, 345, 939. Burrows, A., Hubbard, W. B., Lunine, J. I., and Liebert, J. 2001, Reviews of Modern Physics, 73, 719. Chabrier, G. and Baraffe, I. 2000, ARA&A, 38, 337. Close, L. M., Siegler, N., Freed, M., and Biller, B. 2003, ApJ, 587, 407. Cuntz, M., Saar, S. H., and Musielak, Z. E. 2000, ApJ, 533, L151. Dahn, C. C. et al. 2002, AJ, 124, 1170. Delfosse, X., Tinney, C. G., Forveille, T., Epchtein, N., Borsenberger, J., Fouqué, P., Kimeswenger, S., and Tiphène, D. 1999, A&AS, 135, 41. Dobler, W., Stix, M., and Brandenburg, A. 2004, astro-ph/0410645. Drake, S. A., Simon, T., and Linsky, J. L. 1989, ApJS, 71, 905. Dulk, G. A. and Marsh, K. A. 1982, ApJ, 259, 350. Durney, B. R., De Young, D. S., and Roxburgh, I. W. 1993, Sol. Phys., 145, 207. Fleming, T. A., Giampapa, M. S., Schmitt, J. H. M. M., and Bookbinder, J. A. 1993a, ApJ, 410, 387. Fleming, T. A., Giampapa, M. S., Schmitt, J. H. M. M., and Bookbinder, J. A. 1993b, ApJ, 410, 387. Güdel, M. 2002, ARA&A, 40, 217. Gelino, C. R., Marley, M. S., Holtzman, J. A., Ackerman, A. S., and Lodders, K. 2002, ApJ, 577, 433. Giampapa, M. S., Rosner, R., Kashyap, V., Fleming, T. A., Schmitt, J. H. M. M., and Bookbinder, J. A. 1996, ApJ, 463, 707. Gizis, J. E., Monet, D. G., Reid, I. N., Kirkpatrick, J. D., Liebert, J., and Williams, R. J. 2000, AJ, 120, 1085. Gizis, J. E., Reid, I. N., Knapp, G. R., Liebert, J., Kirkpatrick, J. D., Koerner, D. W., and Burgasser, A. J. 2003, AJ, 125, 3302. Goldreich, P. and Soter, S. 1966, Icarus, 5, 375. Gudel, M., Schmitt, J. H. M. M., Bookbinder, J. A., and Fleming, T. A. 1993, ApJ, 415, 236. Guedel, M. and Benz, A. O. 1993, ApJ, 405, L63. Guedel, M., Benz, A. O., Schmitt, J. H. M. M., and Skinner, S. L. 1996, ApJ, 471, 1002. Haisch, B., Strong, K. T., and Rodono, M. 1991, Ann. Rev. Astr. Ap., 29, 275. Hambaryan, V., Staude, A., Schwope, A. D., Scholz, R.-D., Kimeswenger, S., and Neuhäuser, R. 2004, A&A, 415, 265. Hawley, S. L. et al. 2002, AJ, 123, 3409. Hawley, S. L. et al. 1995, ApJ, 453, 464. Johns-Krull, C. M. and Valenti, J. A. 1996, ApJ, 459, L95+. Kirkpatrick, J. D. et al. 2000, AJ, 120, 447. Leggett, S. K. et al. 2000, ApJ, 536, L35. Leggett, S. K. et al. 2002, ApJ, 564, 452. 11 Leto, G., Pagano, I., Linsky, J. L., Rodonò, M., and Umana, G. 2000, A&A, 359, 1035. Liebert, J., Kirkpatrick, J. D., Cruz, K. L., Reid, I. N., Burgasser, A., Tinney, C. G., and Gizis, J. E. 2003, AJ, 125, 343. Linsky, J. L. and Gary, D. E. 1983, ApJ, 274, 776. Luhman, K. L., Rieke, G. H., Young, E. T., Cotera, A. S., Chen, H., Rieke, M. J., Schneider, G., and Thompson, R. I. 2000, ApJ, 540, 1016. Martín, E. L. and Ardila, D. R. 2001, AJ, 121, 2758. Martín, E. L., Basri, G., Delfosse, X., and Forveille, T. 1997, A&A, 327, L29. Martín, E. L. and Bouy, H. 2002, New Astronomy, 7, 595. Martín, E. L., Delfosse, X., Basri, G., Goldman, B., Forveille, T., and Zapatero Osorio, M. R. 1999, AJ, 118, 2466. Mohanty, S. and Basri, G. 2003, ApJ, 583, 451. Mohanty, S., Basri, G., Shu, F., Allard, F., and Chabrier, G. 2002, ApJ, 571, 469. Mutel, R. L. and Lestrade, J. F. 1985, AJ, 90, 493. Neupert, W. M. 1968, ApJ, 153, L59+. Osten, R., Hawley, S., Allred, J., Johns-Krull, C., and Roark, C. 2004, astro-ph/0411236. Pizzolato, N., Maggio, A., Micela, G., Sciortino, S., and Ventura, P. 2003, A&A, 397, 147. Press, W. H., Teukolsky, S. A., Vetterling, W. T., and Flannery, B. P. 1992, Numerical recipes in FORTRAN. The art of scientific computing, (: Cambridge: University Press, c1992, 2nd ed.). Putman, M. E. and Burgasser, A. J. 2003, American Astronomical Society Meeting, 203, . Raedler, K.-H., Wiedemann, E., Meinel, R., Brandenburg, A., and Tuominen, I. 1990, A&A, 239, 413. Reid, I. N., Gizis, J. E., Kirkpatrick, J. D., and Koerner, D. W. 2001, AJ, 121, 489. Reid, I. N., Kirkpatrick, J. D., Gizis, J. E., Dahn, C. C., Monet, D. G., Williams, R. J., Liebert, J., and Burgasser, A. J. 2000, AJ, 119, 369. Reid, I. N., Kirkpatrick, J. D., Gizis, J. E., and Liebert, J. 1999a, ApJ, 527, L105. Reid, I. N. et al. 1999b, ApJ, 521, 613. Roberts, D. H., Lehar, J., and Dreher, J. W. 1987, AJ, 93, 968. Rosner, R., Golub, L., and Vaiana, G. S. 1985, ARA&A, 23, 413. Rutledge, R. E., Basri, G., Martín, E. L., and Bildsten, L. 2000, ApJ, 538, L141. Saar, S. H. and Linsky, J. L. 1985, ApJ, 299, L47. Schweitzer, A., Gizis, J. E., Hauschildt, P. H., Allard, F., and Reid, I. N. 2001, ApJ, 555, 368. Shkolnik, E., Walker, G. A. H., and Bohlender, D. A. 2003, ApJ, 597, 1092. Smith, E. J., Davis, L., Jones, D. E., Coleman, P. J., Colburn, D. S., Dyal, P., and Sonett, C. P. 1975, Science, 188, 451. Soderblom, D. R., Stauffer, J. R., Hudon, J. D., and Jones, B. F. 1993, ApJS, 85, 315. Stelzer, B. 2004, ApJ, 615, L153. Stepanov, A. V., Kliem, B., Zaitsev, V. V., Fürst, E., Jessner, A., Krüger, A., Hildebrandt, J., and Schmitt, J. H. M. M. 2001, A&A, 374, 1072. Stevenson, D. J. 1991, ARA&A, 29, 163. Vrba, F. J. et al. 2004, AJ, 127, 2948. West, A. A. et al. 2004, AJ, 128, 426. Zapolsky, H. S. and Salpeter, E. E. 1969, ApJ, 158, 809. 12 ) y J (m 6 4 . 8 , Fn ) y J (m 6 8 . 4 , Fn 800 600 400 200 0 −200 −400 −600 −800 800 600 400 200 0 −200 −400 −600 −800 0 Stokes I Stokes V 8.46 GHz 4.86 GHz 1 Time Since 2002 Sep. 28.08 UT (hours) 5 2 3 4 6 7 8 FIG. 1. -- Radio light curves of 2MASS J00361617+1821104 at 8.5 GHz (top) and 4.9 GHz (bottom). Both total intensity (black circles) and circular polarization (gray diamonds) are shown, with a 10-min time resolution. The negative sign of the circularly- polarized flux indicates left-handed polarization. The light curves for both frequencies and polarizations are variable, though with lower significance at 8.5 GHz (§6). 13 Field Sources Stokes I Stokes V 1 Time Since 2002 Sep. 28.08 UT (hours) 5 2 3 4 6 7 8 n o i t i a v e D l a n o i t c a r F n o i t i a v e D l a n o i t c a r F n o i t i a v e D l a n o i t c a r F 3 2 1 0 −1 −2 3 2 1 0 −1 −2 3 2 1 0 −1 −2 0 FIG. 2. -- Fractional deviation relative to the mean of the circular polarization (bottom) and total intensity (middle) light curves of 2MASS J00361617+1821104 at 4.9 GHz. Time resolutions of 10, 15, and 20 minutes are plotted. The top panel shows the light curves of two field sources with a 10-min time resolution. Clearly, the emission from 2M0036+18 is highly variable independent of time resolution. The light curve of 2M0036+18 also appears to be periodic with a period of about 3 hours. 14 1.2 1 0.8 0.6 0.4 0.2 ) 1 − s 2 − m c g r e ( Fl 0 6500 7000 7500 8000 8500 9000 9500 10000 Wave le ng th (A) FIG. 3. -- Combined 4-hour optical spectrum of 2M0036+18. The location of Hα and prominent absorption features are marked. Clearly, no Hα emission is detected above the continuum level (inset). x 10−15 8 6 4 2 0 x 10−17 Ha Li I 6500 6600 6700 6800 TiO CrH Na I Ha Li I CaH Rb I K I Cs I FeH FeH H O 2 15 −11 −12 −13 −14 −15 −16 M0 M5 L0 L5 ) 1 − z H ( ] L / L [ g o l X R ) s e r a l f ( 0 2 − 4 4 9 P L 8 1 + 6 3 0 0 M 2 4 1 2 0 − 1 2 0 0 R B I ) t n e i t s s r e p ( 0 2 − 4 4 9 P L Sun Stars M0−M7 Late−M & L 1018 1017 1016 1015 1014 1013 1012 1011 1010 109 108 107 ) 1 − z H 1 − s g r e ( L R 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 (erg s−1) L X FIG. 4. -- Radio versus X-ray luminosity for stars exhibiting coronal activity. Data for late-M and L dwarfs are from B01 and B02, while other data are taken from Güdel (2002) and references therein. Data points for the Sun include impulsive and gradual flares, as well as microflares. The strong correlation between LR and LX is evident and extends to spectral type M7 (see inset). Clearly, 2MASS J00361617+1821104 and the other late-M and L dwarfs detected in the radio to date violate the correlation by many orders of magnitude. 16 l n o i t a e r r o C − o t u A 1.2 1 0.8 0.6 0.4 0.2 0 0 Stokes I Stokes V 2002 Sep. 28 UT 1 2 3 4 T (hours) 5 6 7 8 FIG. 5. -- Auto-correlation function (ACF) as a function of temporal lag for the total intensity (black circles) and circular polarization (gray diamond) light curves of 2MASS J00361617+1821104, as well as two field sources (gray squares). The time resolution is 10 minutes (thick lines and field sources). A clear peak in the ACF of 2M0036+18 is seen at a lag of 3 hours. The flat ACF of the field sources is indicative of a non-periodic signal. The timescale of about 1 hour on which the signal from 2M0036+18 de-correlates (i.e., the first minimum in the ACF) indicates that the covering fraction of the emission region is ∼ 10%. D 17 99.99% 99.9% 99.7% 95.5% 68.3% d t=7 min d t=8 min 4000 3000 2000 1000 0 4000 3000 2000 1000 0 1 2 3 4 5 6 7 8 9 10 11 Peak Power 2 1.5 Frequency (hr−1) 2.5 3 3.5 4 12 10 8 6 4 2 r e w o P 0 0 0.5 1 FIG. 6. -- Lomb-Scargle periodogram for the circularly polarized emission at 4.9 GHz with a time resolution of 8 min (black) and 7 min (gray). The dashed lines indicate the 68.3%, 95.5%, 99.7%, 99.9%, and 99.99% significance levels based on Monte Carlo simulations of 50,000 randomized versions of the light curve (inset). For both time resolutions the most significant peak (99.9% and 99.99%) is at a frequency of 0.33 hr- 1, or a period of 3.0 hours. We find the same result, though with different significance (see Figure 7), for both total intensity and circular polarization at a wide range of time resolutions. 18 ) % ( e c n a c i f i n g S i 99.999 99.99 99.9 99 90 Circular Polarization Total Intensity 101 Time Resolution (min) FIG. 7. -- Significance of the 3-hour period in the total intensity (black) and circular polarization (gray) 4.9 GHz light curves as a function of time resolution. The significance is determined using Monte Carlo simulated and randomized light curves as described in §6 and shown in Figure 6. 19 Stokes V Stokes I 1 1.5 2 Time Since 2005 Jan. 10.99 UT (hours) 2.5 3 3.5 4 4.5 5 ) y J (m 6 8 . 4 , Fn ) y J (m 6 8 4 . , Fn 400 200 0 −200 −400 −600 600 400 200 0 −200 −400 0 0.5 FIG. 8. -- Follow-up observations of 2M0036+18 at 4.86 GHz taken on 2005, Jan. 10.99 UT for a total of 5 hours. The 3-hour period detected in the initial observations from Sep. 2002 is clearly seen. 20 ) y J (m 6 8 . 4 , Fn ) y J (m 6 8 4 . , Fn 400 200 0 −200 −400 −600 600 400 200 0 −200 −400 0 0.5 Stokes V Stokes I 1 1.5 2 Time Since 2005 Jan. 11.99 UT (hours) 2.5 3 3.5 4 4.5 5 FIG. 9. -- Follow-up observations of 2M0036+18 at 4.86 GHz taken on 2005, Jan. 11.99 UT for a total of 5 hours. The 3-hour period detected in the initial observations from Sep. 2002 is clearly seen. 21 3.5 4 4.5 5 2005 Jan. 10.99 UT 0.5 2005 Jan. 11.99 UT 3 T (hours) 1.5 1 2 2.5 n o i t l a e r r o C − o t u A n o i t l a e r r o C − o t u A 1 0.5 0 1 0 0.5 0 0 0.5 1 1.5 2 2.5 3 T (hours) 3.5 4 4.5 5 FIG. 10. -- Auto-correlation function (ACF) as a function of temporal lag for circular polarization in the follow-up observations on Jan. 10 (top) and Jan. 11 (bottom). The time resolution ranges from 2 to 15 min. A clear peak is seen at a lag of 3 hours, with a fall-off time of about 1 hour, confirming the periodicity. D D 22 ) y J (m y t i s n e D x u F l 600 400 200 0 −200 −400 −600 0 Stokes I Stokes V T = 184 min 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Phase 1 FIG. 11. -- Phased light curve (total intensity and circular polarization) for the observations from Jan. 10 and 11 UT folded with a period of 184 min. 23 1 0.5 0 Rb I l x u F d e z i l a m r o N 1 0.5 0 l x u F d e z i l a m r o N Rb I 7790 7795 7800 7805 7810 7815 7935 7940 7945 7950 7955 7960 Wavelength (A) Wavelength (A) l x u F d e z i l a m r o N 1 0.5 0 l x u F d e z i l a m r o N 1 0.5 0 Na I Cs I 8170 8180 8190 8200 8210 8510 8515 8520 8525 8530 8535 Wavelength (A) Wavelength (A) FIG. 12. -- Zoom-in on four spectral features in the high-resolution MIKE spectra of 2M0036+18, smoothed by two times the instrumental resolution. The second spectrum has been offset downward for clarity. No obvious shift in the lines is discernible. 24 n o i t l a e r r o C − s s o r C −10 −8 −6 −4 n o i t l a e r r o C − s s o r C n o i t l a e r r o C − s s o r C −2 Radial Velocity (km s−1) 0 2 4 6 8 10 Rb I Rb I Na I Cs I −7 −5 −3 −1 0 1 ∆λ (A) 3 5 7 −7 −5 −3 −1 0 1 ∆λ (A) 3 5 7 FIG. 13. -- Cross-correlation of the two MIKE spectra using the orders in which strong absorption lines were detected. The peak is at zero shift indicating no detectable radial velocity signature between the two epochs. The limiting factor in the accuracy of this result is the width of the absorption features. ] l o b R L / L [ g o l −5 −6 −7 −8 −9 −10 ] l o b X L / L [ g o l ] l o b L / Ha L [ g o l −1 −2 −3 −4 −5 −6 −2 −3 −4 −5 −6 −7 25 2M 0036+18 Radio X−ray 2M 0036+18 Ha 2M 0036+18 M2 M3 M4 M5 M6 M7 M8 M9 L0 L1 L2 L3 L4 L5 FIG. 14. -- Radio (top), X-ray (middle) and Hα (bottom) activity as a function of spectral type. Shown are objects with flares (solid squares) and persistent emission (detections: open squares; upper limits: triangles), as well as objects with persistent emission only (detections: circles; upper limits: triangles). The solid line in each panel marks the maximum level of activity for spectral types earlier than M7, and the dashed lines are an extrapolation to later spectral types. In both X-rays and Hα the level of persistent activity drops significantly beyond M7, while flares occasionally reach the same level of activity of early M dwarfs. In the radio band, on the other hand, several objects exhibit activity levels much stronger than that observed in early M dwarfs. This is particularly pronounced in the case of 2M0036+18. 26 ] l o b R L / L [ g o l ] l o b X L / L [ g o l ] l o b L / Ha L [ g o l −5.5 −6 −6.5 −7 −7.5 −8 −8.5 −9 −9.5 −2.5 −3 −3.5 −4 −4.5 −5 −5.5 −6 −6.5 −3.5 −4 −4.5 −5 −5.5 −6 −6.5 −7 −7.5 2M0036+18 2M0036+18 2M0036+18 0 10 20 30 40 vsin i (km s−1) Radio X−ray Ha 50 60 FIG. 15. -- Radio (top), X-ray (middle) and Hα (bottom) activity as a function of rotation velocity. M dwarfs up to spectral type M8 are shown in gray symbols, while dwarfs later than M8 are shown in black symbols. The early M dwarfs exhibit a clear rotation- activity relation in Hα and X-rays (gray dashed lines). This relation breaks down in the late-M and L dwarfs with many fast rotators exhibiting no Hα or X-ray emission. In the radio, on the other hand, faster rotation appears to be correlated with activity even for dwarfs later than M9. The large open circle designates the L5 dwarf 2M 1507-16 (see Appendix).
astro-ph/9809065
1
9809
1998-09-05T16:28:26
Phase Space Transport in Noisy Hamiltonian Systems
[ "astro-ph" ]
This paper analyses the effect of low amplitude friction and noise in accelerating phase space transport in time-independent Hamiltonian systems that exhibit global stochasticity. Numerical experiments reveal that even very weak non-Hamiltonian perturbations can dramatically increase the rate at which an ensemble of orbits penetrates obstructions like cantori or Arnold webs, thus accelerating the approach towards an invariant measure, i.e., a near-microcanonical population of the accessible phase space region. An investigation of first passage times through cantori leads to three conclusions, namely: (i) that, at least for white noise, the detailed form of the perturbation is unimportant, (ii) that the presence or absence of friction is largely irrelevant, and (iii) that, overall, the amplitude of the response to weak noise scales logarithmically in the amplitude of the noise.
astro-ph
astro-ph
to appear in: Annals of the New York Academy of Sciences Phase Space Transport in Noisy Hamiltonian Systems HENRY E. KANDRUPa Department of Astronomy and Department of Physics and Institute for Fundamental Theory University of Florida, Gainesville, Florida 32611 ABSTRACT. This paper analyses the effect of low amplitude friction and noise in acceler- ating phase space transport in time-independent Hamiltonian systems that exhibit global stochasticity. Numerical experiments reveal that even very weak non-Hamiltonian per- turbations can dramatically increase the rate at which an ensemble of orbits penetrates obstructions like cantori or Arnold webs, thus accelerating the approach towards an invari- ant measure, i.e., a microcanonical population of the accessible phase space region. An investigation of first passage times through cantori leads to three conclusions, namely: (i) that, at least for white noise, the detailed form of the perturbation is unimportant, (ii) that the presence or absence of friction is largely irrelevant, and (iii) that, overall, the amplitude of the response to weak noise scales logarithmically in the amplitude of the noise. WHY CONSIDER FRICTION AND NOISE? In general, very weak non-Hamiltonian perturbations will only have very weak effects on properties of flows in time-independent Hamiltonian systems which are integrable or near-integrable and admit no global stochasticity. This also seems to be true for systems completely dominated by chaos where regular orbits are virtually nonexistent. However, low amplitude non-Hamiltonian perturbations can have important qualitative effects on more complex Hamiltonian systems that admit significant measures of both regular and chaotic orbits. For example, weak noise will serve as a source of extrinsic diffusion that can dramatically accelerate phase space transport through cantori (for D = 2) or along Arnold webs (for D ≥ 3). Consider, e.g., flows in two-dimensional systems. Here one knows that, in the absence of friction and noise, cantori [1,2], fractured KAM tori associated with the breakdown of integrability that contain a cantor set of holes, can partition a single connected chaotic phase space region into separate parts which, albeit not completely disjoint, are distinct in the sense that a chaotic orbit starting in one part of the phase space will remain stuck in aHEK was supported in part by National Science Foundation Grant No. PHY92-03333 and by Los Alamos National Laboratory through the Institute of Geophysics and Planetary Physics. Some of the numerical calculations described here were facilitated by computer time provided by IBM through the Northeast Regional Data Center (Florida). 1 that part for a long time before wending its way through one or more holes in the cantori to access another region [3,4]. However, introducing even very weak friction and noise can dramatically increase the rate at which orbits pass through these holes, thus allowing orbits to probe the entire accessible phase space much more quickly. That noise can accelerate phase space diffusion has been long known to dynamicists studying various maps, dating back at least to the work of Lieberman and Lichtenberg [5] in the 1970's. However, the details have not received all that much attention, especially for continuous systems. But why should one care? Why is this phenomenon important in the real world? The crucial observation here is that there is no such thing as a truly isolated system. Every system in nature is coupled to at least some degree to its surrounding environment. The important point then is that, in many cases, one should expect that the coupling of a system to an external environment can be modeled as resulting in friction and noise, related by a Fluctuation-Dissipation Theorem [6,7] (although there are examples where such a picture not justified [8]). Indeed, the assumption of a system coupled by a Fluctuation-Dissipation Theorem to an environment, idealised as a heat bath characterised by some temperature Θ = kBT , is one powerful starting point for modern theories of nonequilibrium statistical mechanics (see, e.g., the textbook by Kubo, Toda, and Hashitsume [9]). Most modeling of this sort involves the assumption of a composite entity of system plus environment which is characterised by a time-independent Hamiltonian H. One might therefore worry that this picture does not extend naturally to a cosmological setting where, in the average comoving frame, the Hamiltonian H typically acquires an explicit time- dependence. Fortunately, however, the assumption of a time-independent H is not essential [10]. Following Caldeira and Legett [6,7], it is natural to write the composite Hamiltonian H as a sum H = Hsys + Hbath + Hint, (1) where the system Hamiltonian Hsys is completely arbitrary, Hbath is idealised as the Hamil- tonian for a collection of linearised excitations, i.e., "phonons," with (in general) time- dependent frequencies, and the interaction Hint is an arbitrary function of the system vari- ables but linear in the bath variables. (These restrictions on Hbath and Hint assure that each mode of the environment is only weakly coupled to the system, so that the environment can be visualised as a thermal "bath.") In this setting, one can always integrate out the explicit dependence on bath variables to derive an exact nonlocal (in time) Langevin equation for the system; and, if Hbath is time-independent, this exact equation (and, presumably, any reasonable Markov approximation thereunto) will satisfy a Fluctuation-Dissipation Theo- rem, regardless of the possible time-dependence of Hsys and Hint. This implies in particu- lar that, for the case of a conformally static Friedmann cosmology, one has a cosmological Fluctuation-Dissipation Theorem whenever the environment can be approximated as a con- formally coupled field, e.g., electromagnetic blackbody radiation.b b In this connection, it should be noted [11] that, in the dipole approximation, the Hamiltonian 2 The natural inference is that, in a variety of different settings, including many relevant to astronomical systems, weak couplings to an external environment will result in non- Hamiltonian perturbations which could have important physical implications. However, the proper form for the noise is not always obvious. The simplest model in terms of which to couple a system to its surroundings, the independent oscillator model [12,13] (which can be shown to be equivalent to many other phenomenological models that have been considered in the past [11]), leads immediately to additive white noise, i.e. state-independent noise that is delta-correlated in time. However, this simple form for the noise is a direct consequence of the assumptions (i) that Hint is linear in the system variables and (ii) that the bath phonons are characterised by an ohmic distribution, i.e., a spectral distribution ∝ ω2dω with appropriate cutoffs. Allowing Hint to involve the system variables in a more complicated fashion leads to multiplicative (i.e., state-dependent) noise; allowing for a different spectral density leads to coloured noise (i.e., noise that is not delta-correlated in time). When considering a galaxy embedded in a rich cluster, one is confronted with a system which, in many cases, is significantly impacted by its surrounding environment. Particularly close encounters between galaxies in a cluster, e.g., those resulting in physical collisions, probably cannot be viewed as "random" events. However, large numbers of relatively weak interactions probably can be viewed as a source of friction and noise, although there is no obvious reason why the noise associated with these interactions can be approximated as delta-correlated in time. Another piece of physics which one might hope to model as friction and noise, again arising in galactic astronomy, is discreteness effects reflecting the fact that a galaxy is com- prised of a collection of nearly point mass stars, rather than the smoothed-out continuum assumed in the context of a description based on the collisionless Boltzmann (i.e., gravita- tional Vlasov) equation. In the context of a smooth one-particle distribution function, these discreteness effects are typically described by a Fokker-Planck, or Landau, equation, which involves a velocity-dependent coefficient of dynamical friction and multiplicative noise (dif- fusion) related by a self-consistent Fluctuation-Dissipation Theorem [14]. Superficially this source of friction and noise might seem completely different from the aforementioned effects associated with an external environment. However this is not really so! In this setting, one can view the full many-particle dynamics as the composite entity of system plus envi- ronment, the reduced one-particle dynamics as the system, and couplings to higher order correlations ignored in a collisionless description as interactions that serve as a source of friction and noise [15]. In the past, a good deal of work has focused on the effects of relatively strong friction and describing the interaction of an electron with a radiation field is, when formulated in an inertial frame, equivalent to the independent oscillator model [12,13], which is perhaps the simplest "realistic" example of the Hamiltonian (1). Transforming to comoving coordinates makes Hsys and Hint time- dependent, but the conformal invariance of the electromagnetic field implies that Hbath remains time-independent. 3 noise in triggering barrier penetration and other phenomena which proceed on the natural relaxation time tR associated with the system's approach towards thermal equilibrium (see, e.g., [14,16] and numerous references cited therein.). This is not the problem of interest here. Rather, the objective of the work described in this paper has been to focus on much weaker perturbations, where tR is much longer than any time scale of interest, and to determine the extent to which friction and noise have significant statistical effects on the evolution of ensembles of orbits already on time scales ≪ tR. The numerical experiments described in the next two sections were performed with the aim of answering three basic questions: 1. How should one visualise the effects of accelerated phase space transport induced by friction and noise? 2. How does the size of the effect scale with the amplitude of the perturbation? 3. To what extent do the details of the perturbation matter? One knows, for example, that multiplicative noise can drive a system towards thermal equilibrium much more quickly than additive noise [17], and, as such, it would seem natural to ask whether multiplicative noise can also accelerate diffusion through cantori and Arnold webs more than additive noise. INVARIANT AND NEAR-INVARIANT DISTRIBUTIONS The computations described in this paper involved integrating Langevin equations of the form dx dt = v and dv dt = −∇Φ − ηv + F, (2) these corresponding to motion in a time-independent Hamiltonian H = v2/2 + Φ(r) which is perturbed by friction and noise. The quantity η represents a coefficient of dynamical friction which, in general, can be a nontrivial function of both r and v. The quantity F is a "random" force, idealised as Gaussian white noise, which is characterised completely by the statistical properties of its first two moments. Specifically, hFi(t)i = 0 and hFi(t1)Fj(t2)i = 2ηΘδijδD(t1 − t2), (3) where i and j label vector components and angular brackets denote an ensemble average. The first of these conditions ensures that the average force vanishes identically. The second ensures that the autocorrelation function is delta-correlated in both direction and time. The normalisation in eq. (3) imposes a Fluctuation-Dissipation Theorem which ensures that, for t → ∞, an arbitrary ensemble of orbits evolved with eqs. (2) will approach a canonical distribution with temperature Θ. 4 To date, integrations have focused on three specific two-dimensional potentials, namely the sixth order truncation of the Toda lattice potential [18], Φ(x, y) = 1 2(cid:16)x2 + y2(cid:17) + x2y − 1 3 y3 + 1 2 x4 + x2y2 + 1 2 y4 + x4y + 2 3 x2y3 − 1 3 y5+ 1 5 x6 + x4y2 + 1 3 x2y4 + 11 45 y6, the so-called dihedral potential [19] for one particular set of parameter values, i.e., Φ(x, y) = −(x2 + y2) + 1 4 (x2 + y2)2 − 1 4 x2y2, and the sum of isotropic and anisotropic Plummer potentials [20] for specified core radii and anisotropy parameters, i.e., (4) (5) (6) V (x, y) = − 1 m (cid:16)c2 + x2 + y2(cid:17)1/2 − (cid:16)c2 + x2 + ay2(cid:17)1/2 , with c = 202/3 ≈ 0.136, a = 0.1 and m = 0.3. In all three cases, the constants were so chosen that, in absolute units, a characteristic crossing time tcr ∼ 1. This implies that, if one visualises these potentials as representing large galaxies like the Milky Way, the Hubble time tH ∼ 100 − 200. As discussed more carefully elsewhere [20], these three potentials manifest very different symmetries. Indeed, the only obvious feature which they share is that, for a variety of energies, they admit significant measures of both regular and chaotic orbits, so that the chaotic phase space regions are significantly impacted by cantori. The fact that, neverthe- less, similar qualitative results were obtained for orbits evolved in all three potentials can thus be interpreted as evidence that the basic conclusions are probably robust. Because all three potentials yielded similar results, the largest number of calculations were performed for the dihedral potential, the most inexpensive computationally, which corresponds physically to a slightly "squared" Mexican hat potential. In all cases, the orbits were computed using a fourth order Runge-Kutta algorithm, noise being implemented using an algorithm developed by Griner et al [21]. Most of the integrations were performed using a time step δt = 10−3. It was verified that a shorter time step δt = 10−4 does not yield significantly different results. The first class of experiments to be performed involved tracking the evolution of ensem- bles of chaotic initial conditions of fixed energy E, selected from some small phase space region in the center of the stochastic sea far from any important cantori. These initial condi- tions were first evolved into the future in the absence of any friction or noise by integrating the deterministic Hamilton equations. They were then reintegrated allowing for friction and noise of variable amplitude. All these experiments assumed additive white noise and friction characterised by a constant η, the two quantities being related by a Fluctuation- Dissipation Theorem. The temperature was frozen at a value Θ ∼ E and the amplitude of the perturbing influences was varied by systematically changing the value of η. 5 In the absence of friction and noise, such orbit ensembles exhibit a two-stage evolution [20,22]. The first stage involves a rapid coarse-grained evolution, proceeding exponentially in time, towards a phase space distribution which is near-invariant in the sense that, once achieved, it only exhibits significant changes on a much longer time scale. Basically, this near-invariant distribution corresponds to a distribution characterised by a nearly constant number density in those portions of the constant energy phase space hypersurface that are not blocked by cantori and a near-zero density everywhere else. The second stage involves a much slower evolution towards what appears to be a true invariant distribution, as orbits in the ensemble diffuse through cantori to access phase space regions that were avoided systematically over shorter time scales. This final invariant distribution corresponds to a microcanonical population of the accessible chaotic regions, i.e., a uniform (in canonical coordinates) population of those portions of the phase space that are accessible to orbits with the specified initial conditions. The time scale for the first stage of the evolution is set at least approximately by the largest short time Lyapunov exponents for the orbits in the ensemble, which determine how fast the ensemble will disperse. Typically this time ∼ tcr. The time scale for the second stage is set by the time scale on which orbits diffuse through cantori, typically ≫ tcr. Suppose now that the orbits are perturbed by friction and noise with η ∼ 10−9 − 10−4, the limiting values here corresponding, respectively, to the typical amplitude associated with discreteness effects in very large and very small galaxies [23,24,25]. In this case, one finds that the time scale associated with the first stage of the evolution is essentially un- changed, i.e., ensembles still approach a near-invariant distribution on a time scale ∼ tcr, but that the time scale for the second stage decreases dramatically! In the absence of fric- tion and noise, the time scale associated with diffusion through cantori typically satisfies tdif f (η = 0) ∼ 103 − 105tcr, but even very weak friction and noise can decrease tdif f (η) by orders of magnitude. For example, η ∼ 10−9 − 10−6 can result in a diffusion time as short as ∼ 100tcr, an interval which, for large galaxies, corresponds to the Hubble time tH. Indeed, for values of η as large as η ∼ 10−4, the diffusion time scale tdif f is often so short that one cannot clearly distinguish between two different stages of evolution. For values of η that large, noisy ensembles exhibit a rapid approach towards a near-invariant distribution that differs significantly from the near-invariant distribution associated with a purely Hamiltonian evolution but is comparatively similar to the true invariant distribu- tion associated with a Hamiltonian evolution. Grey-scale plots comparing representative deterministic and noisy near-invariant distributions in the dihedral and truncated Toda potentials are exhibited, respectively, in FIGURES 8 in [25] and FIGURES 2 in [24]. In the absence of friction and noise, one anticipates that a generic ensemble of ini- tial conditions will ultimately evolve towards a microcanonical distribution, i.e., a uniform population of the accessible portions of the constant energy hypersurface, but this will only happen on the relatively long time scale tdif f (η = 0). Alternatively, if one allows for friction and noise and integrates for a time ∼ tR, one anticipates an evolution towards a canonical 6 distribution with temperature Θ. Noisy integrations performed for a time ≪ tR but still much longer than the time required to breech cantori will result in a near-invariant distribu- tion that can be reasonably visualised as a slightly "thickened" version of a constant energy microcanonical distribution. Because E is not exactly conserved, this near-invariant distri- bution is not exactly microcanonical, i.e., not proportional to a delta function in energy. However, because E is almost conserved and the orbits have succeeded in breeching cantori, this noisy near-invariant distribution is much closer to the purely Hamiltonian invariant dis- tribution than to either a canonical distribution or the purely Hamiltonian near-invariant distribution [25]. In this sense, weak friction and noise can accelerate an approach towards a near-microcanonical equilibrium. FIRST PASSAGE TIME EXPERIMENTS To quantify the rate at which individual trajectories diffuse through cantori, a collection of first passage time experiments was also performed. These involved four components: 1. Select individual initial conditions corresponding in the absence of friction and noise to sticky or confined chaotic orbits, i.e., chaotic orbits which, because of cantori, are trapped near regular regions for relatively long times. 2. Specify the form and amplitude of the friction and noise. 3. For each choice of form and amplitude, perform a large number (∼ 2000 - 5000) of different noisy realisations of the same initial condition; and, for each noisy realisation, determine the time at which the orbit escapes through one or more cantori to become unconfined. 4. Analyse the data to extract N (t), the fraction of the orbits that have not yet escaped within a time t. The results quoted below involve orbits in the dihedral potential (5) with E = 10 where, in the absence of any friction or noise, the diffusion time tdif f ∼ 1000. Other choices of potential or energy can yield results that differ quantitatively, but the principal qualitative conclusions seem unchanged. Estimating when an orbit has escaped was done by identifying a "masked" region in the configuration space and recording the first time that the orbit left this region. That this mask criterion is reasonable was tested in two ways: (1) It was verified that changing slightly the shape and location of the mask had no appreciable effects. (2) For the case of purely Hamiltonian trajectories, escape from the masked region was shown to correspond to an abrupt increase in the value of the largest short time Lyapunov exponent. This is in accord with the fact that, albeit still chaotic, confined chaotic orbits are less unstable exponentially than are unconfined chaotic orbits [20]. One interesting variant of the preceding, also considered, involved tracking a localised ensemble of initial conditions corresponding to confined chaotic orbits evolved into the future both with and without friction and noise. These experiments yielded results very similar to those obtained from multiple integrations of individual initial conditions. Six different forms of friction/noise were considered, namely: (1) additive white noise 7 and a constant coefficient of dynamical friction η, related by a Fluctuation-Dissipation Theorem at temperature Θ = E = 10; (2) multiplicative white noise and dynamical fric- tion with η = η0v2, related by a Fluctuation-Dissipation Theorem with Θ = E = 10; (3) multiplicative white noise and dynamical friction with η = η0v−2, again related by a Fluctuation-Dissipation Theorem with Θ = E = 10; and (4) - (6) the same noises as in (1) - (3) but vanishing friction. In all six cases, the individual noisy realisations were generated using the same pseudo-random seeds. In analysing the effects of friction and noise, attention focused on three principal issues, namely: 1. What is the functional form of N (t), the fraction of the orbits that have not yet escaped? 2. How does N (t) depend on the amplitude of the perturbation? 3. To what extent does the form of the friction and the noise actually matter? Overall, in these experiments escape is a two-stage process. Early on, there are no escapes. All that one sees is that, as one might expect [25,26], different noisy realisations of the same initial condition diverge exponentially at a rate set by the value of the largest short time Lyapunov exponent for the unperturbed deterministic trajectory. Eventually, however, once the noisy ensemble has dispersed to the extent that the root mean squared δrrms ∼ 1.0, individual noisy orbits begin to escape through holes in the cantori. This onset of escape is a comparatively abrupt phenomenon, the interval during which the first 5% of the orbits escape typically being only a small fraction of the time T before the first escape occurs. It is also clear that, at least early on, escapes can be well approximated as a Poisson process, with the confined orbits becoming unconfined at a nearly constant rate, i.e., N (t) ≈ ( N (0), N (0) exp [−Λ(t − T )], if t ≤ T ; if t > T . (7) This behaviour is illustrated in FIGURES 1 (a) and (b), which exhibit ln N (t) as a function of time t for two different initial conditions integrated in the presence of a constant η and additive white noise. Each panel summarises multiple noisy realisations of a single initial condition evolved for a total time t = 1024 with Θ = 10. The six different curves in each panel, each summarising 4000 noisy reasliations, represent six different values of η, namely ln η = −9, −8, −7, −6, −5, and −4. It is clear from FIGURE 1 that, although ln N (t) originally decreases linearly in time, it eventually develops nontrivial curvature indicating that the escape rate is slowly decreasing. Exactly why this is so is not completely clear. However, two important points should be noted. (1) In every case where it is observed, this curvature arises at a time sufficiently late that changes in energy δE have become appreciable, ∼ 10% or more. This suggests strongly that, at least in part, this change in escape rate reflects changes in the "effective" Hamiltonian phase space in which the noisy orbits evolve. (2) In at least some cases, the curvature reflects the fact that, because of the perturbations, some originally chaotic orbits have become trapped by KAM tori. (If, for the nonescapers, the friction and noise are 8 turned off at some time τ and the trajectories integrated for a significantly longer time, it becomes apparent that many of the orbits have become regular!). So how, overall, does the form of N (t) scale with η, the amplitude of the perturbation? When probing the time T required before escapes begin or the rate Λ at which escapes initially proceed after they begin, one finds a roughly logarithmic dependence on η. In other words, when plotting T (η) or Λ(η) the natural independent variable, i.e., the abscissa, is ln η, not η. An example thereof is provided in FIGURES 2 a and b, which summarise data generated from a single initial condition with friction and additive white noise related by a Fluctuation-Dissipation Theorem. The top panel exhibits T (0.01), the time required for 1% of the orbits in a 4000 orbit ensemble to become unconfined. The lower panel exhibits the best fit value of the escape rate Λ of eq. (7), as fit to the interval T (0.01) < t < 256. The curvature observed in both panels is statistically significant, so one cannot assert that T or Λ are linear functions of ln η. However, it is clear that, overall, T and Λ should be visualised as functions of ln η rather than η. Perhaps the most important conclusion derived from these experiments is that the com- puted N (t) is nearly independent of the presence or absence of friction, and that N (t) is also largely independent of whether the noise is additive or multiplicative! First perform 4000 noisy realisations of the same initial condition, all with the same Θ and the same η(v), and analyse the resulting data to extract N (t). Then repeat these experiments with exactly the same noise (generated from the same pseudo-random seeds!) but without friction, and once again compute N (t). A comparison of the two N (t)'s then shows virtually no appreciable differences. At early times, there are absolutely no statisti- cally significant differences. Later on, one can see some tiny differences. However, these can be attributed entirely to the fact that the energies of the orbits with and without friction will be slightly different, and that slightly different energies can give rise to slightly different escape statistics. Comparing additive and multiplicative noise is a bit more subtle since one must worry about normalisations. Suppose, however, that, when introducing multiplicative noise, one selects η0 so that the "average" η ≡ η0hv2i or η ≡ η0hv−2i coincides with the white noise constant η. In this case, one finds that the form of the noise matters very little. Plots of N (t) for additive white noise, multiplicative noise ∝ v2, and multiplicative noise ∝ v−2 yield no statistically significant differences. Two examples of this behaviour are exhibited in FIGURES 3 (a) and (b) which compare the effects of additive and multiplicative noise for two different initial conditions at two different perturbation levels. Each panel contains four curves, representing (1) additive white noise and friction with a constant η, (2) the same additive white noise with vanishing friction, (3) multiplicative noise and friction with η ∝ v2, and (4) multiplicative noise and friction with η ∝ v−2. It is evident that, at late times, the curves do not completely overlap. However, it is also clear that none of the curves is extremely different from the others. These numerical experiments suggest two obvious inferences which, however, remain to 9 be checked more carefully for larger orbit ensembles, different frictions and noise, and other potentials: (1) Smooth non-Hamiltonian perturbations like friction play only a minimal role in accelerating phase space transport through cantori. (2) At least assuming that the noise is white, its details seem comparatively unimportant. In particular, additive noise and multiplicative noise depending on the orbital velocity v exhibit only minimal differences. Overall, when perturbing the Hamiltonian trajectories what seems important is that the orbits be subjected to highly "irregular" perturbations that violate Liouville's Theorem at some given amplitude. In this context, it should perhaps be noted explicitly that a rapidly varying time-dependent Hamiltonian need not be as efficient in triggering accelerated phase space transport as random noise. Specifically, when considering a Hamiltonian of the form H = H0 + ǫH1(t), with H1(t) periodic in time and ǫ an adjustable parameter, one finds in at least some cases [27] that, for periods ≪ tcr, one must allow for relatively large values of ǫ to dramatically accelerate diffusion through cantori. WORK IN PROGRESS AND POTENTIAL IMPLICATIONS The experiments described above still need to be generalised in two important ways. One obvious tack involves extending the computations to three-dimensional systems. Arnold webs can serve as partial phase space obstructions in the same sense as can can- tori; and one might anticipate that friction and noise could accelerate phase space transport through such barriers in three-dimensional systems in the same ways as they do through can- tori in two dimensions. This remains, however, to be checked. Indeed, for orbit ensembles in three-dimensional Hamiltonian systems, even the purely deterministic evolution towards an an invariant or near-invariant distribution is not completely understood. Preliminary investigations performed by Merritt and Valluri [28] suggest that the approach towards equilibrium can closely resemble what is observed in two-dimensional systems [20,22]. How- ever, one would anticipate that, for a generic three-dimensional system with two positive Lyapunov exponents, the situation could be more complicated -- and interesting -- than in two dimensions since unequal Lyapunov exponents could induce "mixing" that proceeds in different directions at different rates [29]! Another equally important objective is to allow for the effects of coloured noise, where the autocorrelation function hFi(t1)Fj(t2)i is not delta-correlated in time. The assumption of singular, delta-correlated noise is an idealisation never exactly realised in nature, even when modeling high frequency phenomena like discreteness effects in systems interacting via short range forces; and the assumption seems especially unreasonable when trying to model objects like galaxies embedded in a dense cluster environment where individual "random" interactions would seem characterised dimensionally by a time scale ∼ tcr or even larger! Allowing for coloured noise could also be important by providing some insights into the question of exactly how and why non-Hamiltonian perturbations result in accelerated phase space transport. Diffusion through cantori, either in the absence or presence of noise, must 10 be related to some "natural" microscopic time scale(s), but the precise nature of these time scales has not yet been established. Systematically increasing the autocorrelation time from zero (white noise) to values substantially larger will allow one to determine the point at which a finite correlation time actually begins to matter. In summary, the numerical experiments described in this paper lead to at least three tentative conclusions: 1. At least for systems that admit a coexistence of regular and chaotic behaviour, even weak couplings to an external environment, modeled as friction and noise, can dramatically accelerate evolution towards a (near-)microcanonical equilibrium. This suggest in particular that idealising a complex Hamiltonian system as a completely isolated entity may be a very bad idea. 2. There is reason to think that the detailed form of the friction and noise is comparatively unimportant. When assessing the effects of friction and noise in accelerating phase space transport, all that really matters may be the amplitude of the perturbation. If true, this would suggest that it may not be all that hard to satisfactorily model the coupling of one's system to its surrounding environment. The details which are hard to determine may not be very important! 3. "Collisionality," i.e., discreteness effects, may be significantly more important in galactic dynamics than generally recognised. For example, such graininess could serve to destabilise quasi-equilibria which use confined chaotic orbits to support interesting structures such as bars [30] or triaxial cusps [31]. ACKNOWLEDGMENTS It is a pleasure to acknowledge useful interactions with my collaborators, Katja Linden- berg, Elaine Mahon, Ilya Pogorelov, and, especially, Salman Habib. I am also grateful to James Meiss and Donald Lynden-Bell for useful comments and critiques. The final draft of this manuscript was written at the Aspen Center for Physics, the hospitality of which I acknowledge gratefully. 11 REFERENCES 1. AUBRY, S. & G. ANDRE. 1978. In Solitons and Condensed Matter Physics. A. R. Bishop and T. Schneider, Eds.: 264. Springer, Berlin. 2. MATHER, J. N. 1982. Topology 21: 457. 3. MACKAY, R. S., J. D. MEISS, & I. C. PERCIVAL. 1984. Phys. Rev. Lett. 52: 697. 4. MACKAY, R. S., J. D. MEISS, & I. C. PERCIVAL. 1984. Physica D 13: 55. 5. LIEBERMAN, M. A. & A. J. LICHTENBERG. 1972. Phys. Rev. A 5: 1852. 6. CALDEIRA, A. O. & A. J. LEGETT. 1983. Physica A 121: 587. 7. CALDEIRA, A. O. & A. J. LEGETT. 1983. Ann. Phys. (NY) 149: 374. 8. BARONE, P. M. V. B. & A. O. CALDEIRA. 1991. Phys. Rev. A 43: 57. 9. KUBO, R., M. TODA, & N. HASHITSUME. 1991. Statistical Physics II: Nonequilibrium Statistical Mechanics. Springer, Berlin, 2nd edition. 10. HABIB, S. & H. E. KANDRUP. 1992. Phys. Rev. D 46: 5303. 11. FORD, G. W., J. T. LEWIS, & R. F. O'CONNELL. 1988. Phys. Rev. A 37: 4419. 12. FORD, G. W., M. KAC, & P. MAZUR. 1965. J. Math. Phys. 6: 504. 13. ZWANZIG, R. 1973. J. Stat. Phys. 9: 215. 14. CHANDRASEHHAR, S. 1943. Rev. Mod. Phys. 15: 1. 15. KANDRUP, H. E. 1989. Comments on Astrophys. 13: 325. 16. HONERKAMP, J. 1994. Stochastic Dynamics Systems. VCH Publishers, New York. 17. LINDENBERG, K. & V. SESHADRI. 1981. Physica A 109: 481. 18. TODA, M. 1967. J. Phys. Soc. Japan 22: 431. 19. ARMBRUSTER, D., J. GUCKENHEIMER, & S. KIM. 1989. Phys. Lett. A 140: 416. 20. MAHON, M. E., R. A. ABERNATHY, B. O. BRADLEY, & H. E. KANDRUP. 1995. Mon. Not. R. Astr. Soc. 275: 443. 21. GRINER, A., W. STRITTMATTER, & J. HONERKAMP. 1988. J. Stat. Phys. 51: 95. 22. KANDRUP, H. E. & M. E. MAHON. 1994. Phys. Rev. E 49: 3735. 23. KANDRUP, H. E. & M. E. MAHON. 1995. Ann. N. Y. Acad. Sci. 751: 93. 24. HABIB, S., H. E. KANDRUP, & M. E. MAHON. 1996. Phys. Rev. E 53: 5473. 25. HABIB, S., H. E. KANDRUP, & M. E. MAHON. 1997. Astrophys. J. 480: 155. 26. KANDRUP, H. E. & D. E. WILLMES. 1994. Astron. Astrophys. 283: 59. 27. KANDRUP, H. E., R. A. ABERNATHY, & B. O. BRADLEY. 1995. Phys. Rev. E 51: 5287. 28. MERRITT, D. & M. VALLURI. 1996. Astrophys. J. 471: 82. 29. KANDRUP, H. E. 1998. Mon. Not. R. Astr. Soc., submitted. 30. WOZNIAK, H. 1993. In Ergodic Concepts in Stellar Dynamics. V. G. Gurzadyan and D. Pfenniger, Ed. Springer, Berlin. 31. MERRITT, D. & T. FRIDMAN. 1996. Astrophys. J. 460: 136. 12 FIGURE CAPTIONS FIGS. 1 (a) N (t), the fraction of confined chaotic orbits that have not yet escaped to become unconfined, for ensembles of 4000 noisy realisations of the same initial condition evolved in the dihedral potential with E = 10.0, x = 0.0, y = 1.3, vy = 1.75, and vx = vx(x, y, vy, E) > 0. Each orbit was subjected to additive white noise and friction with constant η, related by a Fluctuation-Dissipation Theorem with temperature Θ = 10. Passing from top to bottom at small t, the six curves represent ensembles with η = 10−9, 10−8, 10−7, 10−6, 10−5, and 10−4. (b) The same quantities generated for a different initial condition, namely E = 10.0, x = 0.0, y = 2.7, and vy = 2.25. FIGS. 2 (a) T (0.01), the time required for 1% of the members of an ensemble of 4000 noisy realisations of the same unconfined chaotic orbit with E = 10.0, x = 0.0, y = 1.3, vy = 1.75, and vx = vx(x, y, vy, E) > 0 to become unconfined. Each orbit was subjected to additive white noise and friction with constant η, related by a Fluctuation-Dissipation Theorem with Θ = 10. (b) Λ, the rate at which orbits in the ensemble escape, fit to the interval T (0.01) < t < 256. FIGS. 3 (a) N (t), the fraction of confined chaotic orbits that have not yet escaped to become unconfined, for ensembles of 4000 noisy realisations of the same initial condition evolved in the dihedral potential with E = 10.0, x = 0.0, y = 1.1, vy = 3.35, and vx = vx(x, y, vy, E) > 0. Each orbit was evolved with Θ = 10 and η0 = 10−5. The four curves represent additive white noise and a constant η (solid line), additive noise but no friction (dashed), multiplicative noise and friction with η ∝ v2 (dot-dashed), and multiplicative noise and friction with η ∝ v2 (triple-dot-dashed). (b) The same quantities generated for a different initial condition, namely E = 10.0, x = 0.0, y = 1.3, vy = 1.75, and vx = vx(x, y, vy, E) > 0, now allowing for Θ = 10 and η0 = 10−7. 13 Figure 1. Figure 2. Figure 3.
astro-ph/0402532
1
0402
2004-02-23T18:58:44
Growth of Intermediate-Mass Black Holes in Globular Clusters
[ "astro-ph" ]
We present results of numerical simulations of sequences of binary-single scattering events of black holes in dense stellar environments. The simulations cover a wide range of mass ratios from equal mass objects to 1000:10:10 solar masses and compare purely Newtonian simulations to simulations in which Newtonian encounters are interspersed with gravitational wave emission from the binary. In both cases, the sequence is terminated when the binary's merger time due to gravitational radiation is less than the arrival time of the next interloper. We find that black hole binaries typically merge with a very high eccentricity (0.93 < e < 0.95 pure Newtonian; 0.85 < e < 0.90 with gravitational wave emission) and that adding gravitational wave emission decreases the time to harden a binary until merger by ~ 30% to 40%. We discuss the implications of this work for the formation of intermediate-mass black holes and gravitational wave detection.
astro-ph
astro-ph
Growth of Intermediate-Mass Black Holes in Globular Clusters Kayhan Gultekin M. Coleman Miller Douglas P. Hamilton University of Maryland, College Park, Dept. of Astronomy ABSTRACT We present results of numerical simulations of sequences of binary-single scat- tering events of black holes in dense stellar environments. The simulations cover a wide range of mass ratios from equal mass objects to 1000:10:10 M⊙ and com- pare purely Newtonian simulations to simulations in which Newtonian encounters are interspersed with gravitational wave emission from the binary. In both cases, the sequence is terminated when the binary's merger time due to gravitational radiation is less than the arrival time of the next interloper. We find that black hole binaries typically merge with a very high eccentricity (0.93 ≤ e ≤ 0.95 pure Newtonian; 0.85 ≤ e ≤ 0.90 with gravitational wave emission) and that adding gravitational wave emission decreases the time to harden a binary until merger by ∼ 30 to 40%. We discuss the implications of this work for the formation of intermediate-mass black holes and gravitational wave detection. Subject headings: black hole physics -- galaxies: star clusters -- globular clusters: general -- stellar dynamics 4 0 0 2 b e F 3 2 1 v 2 3 5 2 0 4 0 / h p - o r t s a : v i X r a 1. Introduction Recent observations suggest that large black holes may reside in the centers of some stellar clusters. X-ray observations in the last few years have shown unresolved sources in galaxies offset from their nuclei and with fluxes that, if isotropic, correspond to luminosities of L ≈ 1039 to 1041 erg s−1 (e.g., Fabbiano, Schweizer, & Mackie 1997; Colbert & Mushotzky 1999; Matsumoto et al. 2001; Fabbiano, Zezas, & Murray 2001). Many of these sources are associated with stellar clusters (Fabbiano et al. 1997; Angelini, Loewenstein, & Mushotzky 2001). The strong variability observed in these sources suggests that they are black holes, and if the observed fluxes are neither strongly beamed nor super-Eddington, the implied -- 2 -- masses are as high as M & 103 M⊙. The fact the sources are non-nuclear implies masses M . 106 M⊙ since a larger mass would have rapidly sunk to the center of the host galaxy due to dynamical friction (< 109 yr for a dispersion velocity of 100 km s−1 and a separation from the galaxy nucleus of 102 pc as in the case of M82; Kaaret et al. 2001). In addition, optical observations of the globular clusters M15 and G1 show velocity profiles consistent with central black holes with masses of 2.5 × 103 M⊙ and 2.0 × 104 M⊙, respectively (Gebhardt et al. 2000; Gerssen et al. 2002; van der Marel, et al. 2002; Gebhardt, Rich, & Ho 2002), although Baumgardt et al. (2003) demonstrate with their N-body simulations that the observations of G1 can be explained without a large black hole. Such intermediate-mass black holes (IMBHs) would be in a different mass category, and thus likely indicative of a different formation scenario, from either 3 - 20 M⊙ stellar-mass black holes, which are thought to be the result of core-collapse supernovae, or 106 - 1010 M⊙ supermassive black holes, which are found in the centers of many galaxies. Several models have been proposed to account for the origin of IMBHs. Madau & Rees (2001) and Schneider et al. (2002) suggest that they are the remnants of massive (M & 200 M⊙) Population III stars. The low metallicity of these stars precludes cooling through metal line emission and enables them to reach masses much larger than ordinary main sequence stars. These large stars avoid significant mass loss due to stellar winds or pulsations, and the star may collapse to form a black hole with almost the same mass as the progenitor star. Portegies Zwart & McMillan (2002) and Gurkan, Freitag, & Rasio (2004) show with numerical simulations that the core of a young stellar cluster may collapse rapidly such that direct collisions of stars will lead to runaway growth of a single object with as much as 10−3 of the original cluster mass over the course of a few million years. Miller & Hamilton (2002a) propose that over a Hubble time stellar-mass black holes in dense globular clusters may grow by mergers to the inferred IMBH masses. In their model, a black hole with mass greater than 50 M⊙ will interact with other massive objects to form binaries that will merge due to gravitational radiation. The merger process may proceed more quickly in the presence of encounters with a third black hole or another black hole binary (Miller & Hamilton 2002b) if the encounters shrink the binary's orbit, as is known to happen with hard (tight) binaries (Heggie 1975). Wherever and however IMBHs formed, the best candidates are found in stellar clusters where three-body encounters are important. An IMBH in a cluster, whether formed there or later swallowed by the cluster, will find its way to the center. As all of the heaviest objects in a cluster sink to the center in a process known as mass segregation, the IMBH will interact primarily with other massive objects and binaries (Sigurdsson & Hernquist 1993; Fregeau et al. 2002). A single IMBH will tend to acquire companions through exchanges with binaries because the most massive pair of objects in a three-body encounter preferentially end up in -- 3 -- the binary (e.g., Heggie, Hut, & McMillan 1996). The IMBH binary will encounter other objects in the dense center of its host cluster, harden further, and ultimately merge. These events are important sources of gravitational waves. The Advanced LIGO (Laser Interferometer Gravitational Wave Observatory) detector is expected to be capable of de- tecting mergers of IMBHs with M . 100 M⊙ (Barish 2000), and LISA (Laser Interferometer Space Antenna) is expected to detect the earlier inspiral phase of an IMBH merger (Danz- mann 2000). In order to predict the gravitational wave signature of the inspiral, the expected separations and eccentricities of the binaries must be known. Because three-body encounters alter the orbital parameters, simulations are needed to predict their distributions as well as the source population and event rates. The three-body problem has been studied extensively, but with every new generation of computing power, our understanding of the problem advances with a wider range of numer- ical simulations and a changing perspective on this rich but conceptually simple problem. Previous studies of the three-body problem have tended to focus on the case of equal or nearly equal masses (e.g., Heggie 1975; Hut & Bahcall 1983) though other mass ratios have been studied (e.g., Fullerton & Hills 1982; Sigurdsson & Phinney 1993; Heggie et al. 1996). The nearly equal mass case does not apply to the case of an IMBH in the core of a stellar cluster. In addition the vast majority of previous work has studied the effect of a single en- counter on a binary. To determine the ultimate fate of an IMBH, simulations of sequences of encounters are needed. Furthermore, to our knowledge no previous work has considered the effects of orbital decay due to gravitational radiation between encounters, which we expect to be important for very tight binaries. In this paper we present numerical simulations of sequences of high-mass ratio binary- single encounters. We describe the code used to simulate the encounters in § 2. Next, we present results of the simulations of sequences of encounters on a range of mass ratios with Newtonian gravity (§ 3.1) and with gravitational radiation between encounters (§ 3.2) and show that including gravitational radiation decreases the duration of the sequence by ∼ 30 to 40%. In § 4 and § 5 we discuss the implications of these results for IMBH formation and gravitational wave detection. 2. Numerical Method We perform numerical simulations of the interactions of a massive binary in a stellar cluster. Simulating the full cluster is beyond current N-body techniques, so we focus instead on a sequence of three-body encounters. Massive cluster objects, such as IMBHs and tight -- 4 -- binary systems, tend to sink the centers of clusters so that a single IMBH is very likely to meet a binary (Sigurdsson & Phinney 1995). Exchanges in which the IMBH acquires a close companion are common. Such a binary in a stellar cluster core will experience repeated interactions with additional objects as long as the recoils from these interactions do not eject the binary. Therefore, we simulate a sequence of encounters between a hard binary and an interloper. We perform one interaction and then use the resulting binary for the next encounter. This is repeated multiple times until the binary finally merges due to gravitational radiation. Because typical velocities involved are non-relativistic and the black holes are tiny compared to their separations, they are treated as Newtonian point masses. In order to test the influence of the binary's mass, we use a range of binary mass ratios. To simplify the problem we study a binary with mass ratio of N:10 M⊙ and a 10 M⊙ interloper, designated as N:10:10, and vary N between 10 M⊙ and 103 M⊙. The simulations were done using a binary-single scattering code that was written to be as general purpose as possible. Because of the vast parameter space that needs to be covered, the code uses a Monte Carlo initial condition generator. The orbits are integrated using hnbody, a hierarchical, direct N-body integrator, with the adaptive fourth order Runge Kutta integrator option (K. Rauch & D. Hamilton, in preparation)1. Because we focus on close approaches where a wide range of timescales are important, an adaptive scheme is often better than symplectic methods. In wide hierarchical triples, direct integration can consume a large amount of compu- tational time. To reduce this, we employ a two-body approximation scheme that tracks the phase of the inner binary. For a sufficiently large outer orbit, the orbit is approximately that of an object about the center of mass of the binary. We calculate this approximate two-body orbit analytically and keep track of the inner binary's phase. When the outer object nears the binary again, we revert to direct numerical integration. The orbit is integrated until one of three conditions is met: 1) one mass departs along a hyperbolic path, 2) the system forms a hierarchical triple with outer semimajor axis greater than 2000 AU, an orbit so large that it would likely be perturbed in the high density of a cluster core and not return, or 3) the integration is prohibitively long, in which case the encounter is discarded and restarted with new randomly generated initial conditions. Roughly 10−4 of all encounters had to be restarted with most occurring for higher mass ratios where resonant encounters (encounters that have more than one close approach and are not simple fly-bys) are more common. In half of our simulations, we evolve the binary's orbit due to gravitational wave emission after each encounter. Since a binary in a cluster 1See http://janus.astro.umd.edu/HNBody/. -- 5 -- spends most of its time and emits most of its gravitational radiation while waiting for an encounter rather than during an interaction, we only include gravitational radiation between encounters. To isolate this effect, we run simulations both with and without gravitational radiation. We include gravitational radiation by utilizing orbit-averaged expressions for the change in semimajor axis a and eccentricity e with respect to time (Peters 1964): and da dt = − 64 5 G3m0m1 (m0 + m1) c5a3 (1 − e2)7/2 (cid:18)1 + 73 24 e2 + 37 96 e4(cid:19) de dt = − 304 15 G3m0m1 (m0 + m1) c5a4 (1 − e2)5/2 (cid:18)e + 121 304 e3(cid:19) , (1) (2) where m0 and m1 (m0 ≥ m1) are the gravitational masses of the binary pair. Here G is the gravitational constant, and c is the speed of light. The orbital elements are evolved until the next encounter takes place, at a time that we choose randomly from an exponential distribution with a mean encounter time, hτenci = 1/ hnv∞σi, where n is the number density of objects in the cluster's core, v∞ is the relative velocity, and σ is the cross-section of the binary. If we assume the mass of the binary m0 + m1 ≫ m2, then σ ≈ πr2 p + 4πrpG (m0 + m1) /v2 ∞, (3) where rp is the maximum considered close approach of m2 to the binary's center of mass. For a thermal distribution of stellar speeds, v∞ = (mavg/m2)1/2 vms, where mavg = 0.4 M⊙ is the average mass of the main sequence star and vms is the main sequence velocity dispersion. In our simulations, the second term of Eq. 3, gravitational focusing, dominates over the first. Averaging over velocity (assumed to be Maxwellian) we find hτenci = 2 × 107(cid:16) vms 10 km s−1(cid:17)(cid:18) 106 pc−3 n (cid:19)(cid:18)1 AU rp (cid:19)(cid:18) 1 M⊙ m0 + m1(cid:19)(cid:18)1 M⊙ m2 (cid:19)1/2 yr. (4) We then subject the binary to another encounter using orbital parameters adjusted by both the previous encounter and the gravitational radiation emitted between the encounters. This sequence of encounters continues until the binary merges due to gravitational wave emission. If orbital decay is not being calculated, then we determine that the binary has merged when the randomly drawn encounter time is longer than the timescale to merger, which is approximately τmerge ≈ 6 × 1017 (1 M⊙)3 m0m1 (m0 + m1) (cid:16) a 1 AU(cid:17)4 (cid:0)1 − e2(cid:1)7/2 yr (5) -- 6 -- for the high eccentricities of importance in this paper. Global energy and angular momentum are monitored to ensure accurate integration. The code also keeps track of the duration of encounters, the time between encounters, changes in semimajor axis and eccentricity, and exchanges (events in which the interloping mass replaces one of the original members of the binary and the replaced member escapes). As a test of our code, we compared simulations of several individual three-body encoun- ters to compare with the work of Heggie et al. (1996). As part of a series of works examining binary-single star scattering events, Heggie et al. (1996) performed numerical simulations of very hard binaries with a wide range of mass ratios and calculated their cross-sections for exchange. We ran simulations of one encounter each of a sample of mass ratios for com- parison. To facilitate comparison of encounters with differing masses, semimajor axes, and relative velocities of hard binaries, Heggie et al. (1996) use a dimensionless cross-section, ¯σ = 2v2 ∞Σ πG (m0 + m1 + m2) a , (6) where v∞ is the relative velocity of the interloper and the binary's center of mass at infinity and Σ is the physical cross-section for exchanges. We calculate Σ as the product of the fraction of encounters that result in an exchange (fex) and the total cross-section of encoun- ters considered: fexπb2 max, where bmax is an impact parameter large enough to encompass all exchange reactions. Our cross-sections are in agreement with those of Heggie et al. (1996) within the combined statistical uncertainty as seen in Table 1. 3. Simulations and Results We used our code to run numerical experiments of three-body encounter sequences with a variety of mass ratios. The binaries consisted of a dominant body with mass, m0 = 10, 20, 30, 50, 100, 200, 300, 500, or 1000 M⊙ and a secondary of mass m1 = 10 M⊙. Because of mass segregation, the objects that the binary encounters will be the heaviest objects in the cluster. In order to simplify the problem, we consider only interactions with interlopers of mass m2 = 10 M⊙. The binary starts with a circular a = 10 AU orbit, and the interloper has a relative speed at infinity of v∞ = 10 km s−1 and an impact parameter, b, relative to the center of mass of the binary such that the pericenter distance of the hyperbolic encounter would range from rp = 0 to 5a. For all binaries, vcirc = [G (m0 + m1) /a]1/2 ≥ 40 km s−1 ≫ v∞, and thus all are considered hard. The Monte Carlo initial condition generator distributes the orientations and directions of encounters isotropically in space, and the initial phase of the binary is randomized such that it is distributed equally in time. We assume the cluster -- 7 -- core has a density of n = 105 pc−3 and an escape velocity of vesc = 50 km s−1 for the duration of the simulation. We discuss the consequences of changing the escape velocity in § 4. For each mass ratio, we simulate 1000 sequences with and without gravitational radiation between encounters. 3.1. Pure Newtonian Sequences Figure 1a shows the change of semimajor axis and pericenter distance as a function of time over the course of a typical Newtonian sequence. The encounters themselves take much less time then the period between encounters, so a binary spends virtually all its time waiting for an interloper. Most of the time in this example is spent hardening the orbit from 1 AU to 0.4 AU because as the binary shrinks, its cross-section decreases and the timescale to the next encounter increases. Figure 1b shows the same sequence plotted as a function of number of encounters. The semimajor axis decreases by a roughly constant factor with each encounter. This is expected for a hard binary, which, according to Heggie's Law (Heggie 1975), tends to harden with each encounter at a rate independent of its hardness. The eccentricity and therefore the pericenter distance, rp = a (1 − e), however, can change dramatically in a single encounter (for a discussion on eccentricity change of a binary in a cluster, see Heggie & Rasio 1996). This sequence ends with a very high eccentricity (e = 0.968), which reduces the merger time given by Eq. 5 to less than τenc. Table 2 summarizes our main results and shows a number of interesting trends. The average number of encounters per sequence, hnenci, increases with increasing mass ratio since the energy that the interloper can carry away scales as ∆E/E ∼ m1/ (m0 + m1) (Quinlan 1996) and since nenc ∼ E/∆E for a constant eccentricity. Energy conservation assures that every hardening event results in an increased relative velocity between the binary and the single black hole. If the velocity of the single black hole relative to the barycenter, and thus the globular cluster, is greater than the escape velocity of the cluster core (typically vesc = 50 km s−1 for a dense cluster; see Webbink 1985), then the single mass will be ejected from the cluster. The average number of ejected masses per sequence, hneji, also increases with increasing mass ratio because the higher mass ratio sequences have a larger number of encounters and because the larger mass at a given semimajor axis has more energy for the interloper to tap. Conservation of momentum guarantees that when a mass is ejected from the cluster at very high velocity, the binary may also be ejected. Table 2 lists hfbineji, the fraction of sequences that result in the ejection of the binary from the cluster. As expected, the fraction decreases sharply with increasing mass such that virtually none of the binaries with mass greater than 300 M⊙ escape the cluster. -- 8 -- Fig. 1. -- Newtonian 1000:10:10 sequence. These panels show the semimajor axes (upper lines) and pericenter distances (lower lines) as functions of time (left panel) and number of encounters (right panel) for one sequence of encounters with no gravitational wave emission. Each change in a and rp is the result of a three-body encounter. Since the binary is hard, the semimajor axis gradually tightens by a roughly constant fractional amount per encounter with most of the time spent hardening the final fraction when close encounters are rare. The pericenter distance, however, fluctuates greatly due to large changes in eccentricity during a single encounter. The sequence ends at a very high eccentricity when the binary would merge due to gravitational radiation before the next encounter. -- 9 -- The shape and size of the orbit after its last encounter determine the dominant gravita- tional wave emission during the inspiral and are of particular interest to us. The distribution of pre-merger semimajor axes for all mass ratios is shown in Figure 2. The distributions all have a similar shape that drops off at low a because the binary tends to merge before another encounter can harden it. For large orbits the binary will only merge for a high eccentricity, and thus there is a long tail in the histograms towards high a from encounters that resulted in an extremely high eccentricity. The distributions for lower mass ratios are shifted to smaller a because for a given orbit, a less massive binary will take longer to merge. This can also be seen in the mean final semimajor axis, haf i, in Table 2. Figure 3 shows the distribution of binary eccentricities after the final encounter for one mass ratio. The plot is strongly peaked near e = 1, a property shared by all other mass ratios. This distribution is definitely not thermal, which would have a mean eccentricity heith ≈ 0.7. The high eccentricity before merger results from both the strong dependence of merger time on eccentricity and the fact that the eccentricity can change drastically in a single encounter (see Fig. 1). As the semimajor axis decreases by roughly the same fractional amount in each encounter, the eccentricity increases and decreases by potentially large amounts with each strong encounter. When the eccentricity happens to reach a large value, the binary will merge before the next encounter. Figure 4 shows the eccentricity distribution for all encounters after the first 10 for all 1000 sequences with a mass ratio of 1000:10:10. The distribution is roughly thermal up to high eccentricity where the binaries merge. Thus merger selectively removes high eccentricity binaries from a thermal distribution. 3.2. General Relativistic Binary Evolution The addition of gravitational radiation between Newtonian encounters is expected to alter a sequence since it is an extra source of hardening and since it circularizes the binary. Figure 5 shows a typical sequence for the 1000:10:10 mass ratio including gravitational radiation. Three-body interactions drive the binary's eccentricity up to e = 0.959 and its semimajor axis down to a = 0.713 AU. Then starting at t = 2.2 × 106 yr over the course of about ten interactions that only weakly affect the eccentricity and semimajor axis, gravitational radiation causes the orbit to decay to a = 0.550 AU and e = 0.946 while the pericenter distance remains roughly constant. The corresponding semimajor axis change in the Newtonian only sequences in Figure 1 takes 45 encounters and more than twice as long although one must be careful when comparing two individual sequences. Gravitational waves make the most difference when the pericenter distance is small, which is guaranteed at the end of a sequence, but can also happen in the middle as Figure 5 shows. -- 10 -- Fig. 2. -- Histograms of final semimajor axes for all mass ratios. The solid histograms are pure Newtonian sequences, and the hatched histograms are sequences with gravitational radiation between encounters. The histograms all have similar shapes with a sharp drop at low a since the binary tends to merge before another encounter can harden it, and it has long tail at high a where the binary will only merge with high eccentricity. The sequences with gravitational radiation have falloffs at smaller a than those without due to both the circularization and the extra source of hardening. -- 11 -- Fig. 3. -- Histogram of final eccentricities for 1000:10:10 mass ratio. The solid histogram is from pure Newtonian sequences, and the hatched histogram is from sequences with gravi- tational radiation between encounters. The histogram is cut at e = 0.8 because ef < 0.8 is rare. The histograms have roughly the same shape for both cases and for all mass ratios although the gravitational wave sequences have a consistently lower mean at higher mass ratios because gravitational wave emission damps eccentricities. The histograms show a de- cidedly non-thermal distribution and are strongly peaked near e = 1. Because the timescale to merge due to gravitational radiation is so strongly dependent on e, the binary will merge when it happens to reach a high eccentricity. -- 12 -- Fig. 4. -- Solid line is a histogram of all eccentricities after each encounter except for the first ten for all pure Newtonian sequences of 1000:10:10. The dashed line is a thermal distribution of eccentricities. The distribution is roughly thermal for low eccentricity but deviates for e & 0.6. The expected thermal distribution of eccentricities is altered by losses of high eccentricity orbits to merger. -- 13 -- Fig. 5. -- 1000:10:10 gravitational radiation sequence. Same as Figure 1a but for a sequence with gravitational radiation between encounters. The effects of gravitational radiation can be seen between 2.2 and 2.4 × 106 years. Over this period, the binary undergoes about ten interactions that do not significantly affect its orbit. During this time, the semimajor axis decays from a = 0.713 AU to 0.550 AU while the pericenter distance remains small and roughly constant. When an encounter reduces the eccentricity at 2.4×106 years, gravitational radiation is strongly reduced. Gravitational radiation becomes important again at the end of the sequence. The sequence ends with the binary's merger from gravitational waves. -- 14 -- Table 2 summarizes the effect of adding gravitational radiation. In general the effect is greater at higher masses because gravitational radiation is stronger for a given orbit. Because of the extra energy sink, the binaries merge with fewer encounters, fewer black holes are ejected, and the fraction of sequences in which a binary is ejected is smaller. The most dramatic change is in the duration of the sequence, which gravitational radiation reduces by 27% to 40%. The distributions of final semimajor axes (Fig. 2) and final eccentricities (Fig. 3) have similar shapes to the Newtonian only distributions. Due to the circularizing effect of gravitational radiation, binaries of all mass ratios merge with a smaller hef i than Newtonian only sequences with the largest difference at high mass ratios. Gravitational radiation also produces a smaller haf i for m0 & 300 M⊙. This can be seen in Figure 2 where the gravitational radiation simulations display an excess number of sequences with low af , which is a consequence of the binaries' lower ef . 4. Implications for IMBH Formation We can use these simulations to test the Miller & Hamilton (2002a) model of IMBH formation. We assume that a 50 M⊙ seed black hole with a 10 M⊙ companion will undergo repeated three-body encounters with 10 M⊙ interloping black holes in a globular cluster with vesc = 50 km s−1 and n = 105 pc−3. We also assume that the density of the cluster core remains constant as the IMBH grows. We then test whether the model of Miller & Hamilton (2002a) can build up to IMBH masses, which we take to be 103 M⊙, 1) without ejecting too many black holes from the cluster, 2) without ejecting the IMBH from the cluster, and 3) within the lifetime of the globular cluster. We also test how these depend on escape velocity and seed mass. If the number of black holes ejected is greater than the total number of black holes in the cluster core, then the IMBH cannot build up to the required mass by accreting black holes alone. To calculate the total number of black holes ejected while building up to large masses, we sum the average number of ejections using a linear interpolation of the values in Table 2. Assuming a cluster escape velocity of vesc = 50 km s−1, we find that the total number of black holes ejected when building up to 1000 M⊙ is approximately 6800 for our Newtonian only and 5300 for gravitational radiation simulations. This is far greater than the estimated 102 to 103 black holes available (Portegies Zwart & McMillan 2000). If there were initially one thousand 10 M⊙ black holes in the cluster, mergers of the massive black hole with a series of 10 M⊙ black holes would exhaust half of the black holes in ∼ 2.6 × 108 yr and would ultimately produce a 240 M⊙ black hole. Increasing the seed mass increases the final mass of the IMBH when half of the field black holes run out. If the seed mass were -- 15 -- 100, 200, or 300 M⊙, then the model would produce a 270, 330, or 410 M⊙ black hole after exhausting half of the cluster black hole population in 1.9, 1.1, or 0.8 × 108 yr, respectively. Figure 6 shows the number of black holes ejected as a function of initial black hole mass for a range of escape velocities. Growth times are much shorter than the ∼ 109 yr necessary for stellar-mass black holes to eject each other from the cluster (Sigurdsson & Hernquist 1993; Portegies Zwart & McMillan 2000; J. M. Fregeau, S. A. Rappaport, & V. Corless, in preparation; R. O'leary et al., in preparation). Therefore, self-depletion of stellar-mass black holes is not a limiting factor. Of particular concern is whether the three-body scattering events will eject the binary from the cluster. The black hole can only merge with other black holes while it is in a dense stellar environment. The probability of remaining in the cluster after one sequence is P = 1 − hfbineji. As can be seen in Table 2, once the black hole has built up to ∼ 300 M⊙, it is virtually guaranteed to remain in the cluster. When starting with 50 M⊙, we calculate the total probability of building up to 300 M⊙ to be 0.0356. Figure 7 shows the probability of building up to 300 M⊙ as a function of starting mass for different escape velocities for the gravitational radiation case. Table 3 lists probabilities for selected seed masses and escape velocities for the gravitational radiation case. In a similar manner, we calculate the total time to build up to 1000 M⊙, assuming that the supply of stellar-mass black holes and density remain constant, an assumption which leads to an underestimation of the time. While the time per merger is larger for the smaller masses, the total time is dominated at the higher masses since more mergers are needed for the same fractional increase in mass. For Newtonian only simulations the total time is 1.1 × 109 yr, and for simulations with gravitational radiation the total time is 7.1 × 108 yr. These are much less than the age of the host globular clusters. Figure 8 shows the time to reach a specified mass for both the Newtonian and gravitational radiation cases. Although there is clearly enough time to build IMBHs as Miller & Hamilton (2002a) propose, the issues of whether there are enough stellar-mass black holes and whether the cluster will hold onto the IMBH remain. The combination of an initial mass of 50 M⊙ and an escape velocity of 50 km s−1 is not likely to produce an IMBH in a globular cluster through three-body interactions with 10 M⊙ black holes, but the general process could still produce IMBHs. Miller & Hamilton (2002a) argued that a seed mass of 50 M⊙ would be retained, but for analytical simplicity they assumed that every encounter changed the semimajor axis by the same fractional amount h∆a/ai. Some encounters, however, can decrease the semimajor axis by several times the average value and thus impart much larger kicks. The authors therefore underestimated the minimum initial mass necessary to remain in the cluster. A hierarchical merging of stellar-mass black holes could, however, still produce an IMBH if 1) -- 16 -- Fig. 6. -- Plot of total number of black holes ejected in building up to 1000 M⊙ as a function of seed mass for the gravitational radiation case assuming different escape velocities. The four curves show different assumed cluster escape velocities in km s−1. For all but the largest seed masses, the number of black holes ejected is greater than the estimated ∼ 103 (indicated by the dashed line) present in a young globular cluster. -- 17 -- Fig. 7. -- Plot of an IMBH's probability of remaining in the cluster and building up to 300 M⊙ as a function of starting mass of the dominant black hole for the gravitational radiation case assuming different escape velocities labeled in km s−1. Once the black hole has built up to 300 M⊙ it is very unlikely that it will be ejected from the cluster. The lowest mass binaries are much more readily ejected and thus are very unlikely to survive a sequence of encounters. Miller & Hamilton (2002a) suggest that IMBHs can be built in this manner with a starting mass ≈ 50 M⊙. We find that such small initial masses are likely to be ejected from the cluster core for reasonable escape velocities of dense clusters. -- 18 -- Fig. 8. -- Plot of total time to build up to a certain mass when built by mergers with 10 M⊙ black holes for Newtonian only results and for runs with gravitational radiation between encounters. The Newtonian only simulations are slower to build up, but both cases reach 1000 M⊙ within about 109 years. The time plotted assumes a constant density of black holes for the duration of IMBH formation. -- 19 -- the initial mass of the black hole were were greater than 50 M⊙, 2) the escape velocity of the cluster were greater than 50 km s−1, or 3) additional dynamics were involved. We consider each of these in turn. If the mass of the initial black hole were, e.g., 250 M⊙ before the onset of compact object dynamics, dynamical kicks would not be likely to eject the IMBH, and it would require fewer mergers to reach 1000 M⊙ and thus a smaller population of stellar mass black holes. The initial black hole could start with such a mass if it evolved from a massive Population III star or from a runaway collision of main sequence stars (Portegies Zwart & McMillan 2002; Gurkan et al. 2004), or it could reach such a mass by accretion of young massive stars, which would be torn apart by tidal forces and impart little dynamical kick. If the initial globular cluster mass is high enough (work by Meylan et al. 2001 indicate masses of 107 M⊙ are available), then the cluster's gravity may be strong enough to retain the gas normally expelled by the first generation of supernovae. If that increases the escape velocity to, e.g., vesc = 70 km s−1, the interactions result in a smaller fraction of ejected binaries. The probability of building from 50 M⊙ to 1000 M⊙ then increases by almost an order of magnitude. In addition, processes with lower dynamical kicks could prevent ejection. One promis- ing mechanism is the Kozai resonance (Kozai 1962; Miller & Hamilton 2002b). If a stable hierarchical triple is formed, then resonant processes can pump up the inner binary's ec- centricity high enough so that it would quickly merge due to gravitational radiation and without any dynamical kick to eject the IMBH from the cluster. Two-body captures (cap- tures in which an interloper passes close enough to the isolated IMBH that it becomes bound and merges due to gravitational radiation) would also result in mergers without dy- namical kicks. Both Kozai-resonance-induced mergers and two-body captures are devoid of dynamical kicks, but they would suffer a gravitational radiation recoil. A system in which a 10 M⊙ black hole merges into a 130 M⊙ non-rotating black hole would have a recoil velocity 20 km s−1 ≤ vr ≤ 200 km s−1 (Favata, Hughes, & Holz 2004). Since vr ∼ (m1/m0)2, a merger between a 10 M⊙ black hole and a seed black hole of mass of 250 M⊙, as discussed above, would experience a recoil velocity . 50 km s−1. Mergers with lower mass objects that are torn apart by tidal forces, such as white dwarfs, would receive no gravitational radiation recoil. Finally, a range of interloper masses instead of the simplified single mass population that we used here may also affect retention statistics since a smaller interloper would impart smaller kicks while still contributing to hardening. Increasing the seed mass and the escape velocity will reduce the number of field black holes ejected but not by enough. As seen in Figure 6, using a seed mass m0 = 250 M⊙ and an escape velocity vesc = 70 km s−1 reduces the number of black holes ejected by 40%, but -- 20 -- this is still several factors more than are available. The Kozai-resonance-induced mergers and two-body captures, however, are methods of merging without possibility of ejecting stellar-mass black holes. In order to reach our canonical 1000 M⊙ intermediate mass while ejecting fewer than 103 black holes, 70-80% of the mergers must come from these ejectionless methods. 5. Implications for Gravitational Wave Detection Our simulations make predictions interesting for gravitational wave detection. After the last encounter of a sequence, the binary will merge due to gravitational radiation. As the binary shrinks and circularizes, the frequency of the gravitational radiation emitted passes through the LISA band (10−4 to 100 Hz) (Danzmann 2000) and then through the bands of ground-based detectors such as LIGO, VIRGO, GEO-600, and TAMA (101 to 103 Hz) (Fidecaro et al. 1997; Schilling 1998; Barish 2000; Ando et al. 2002). By the time the binaries are detectable by ground-based instruments, they will have completely circularized, but while in the LISA band, some will have measurable eccentricities. We calculate the distribution of eccentricities detectable by LISA by integrating Equations 1 and 2 until the orbital frequency reaches νorb = 10−3 Hz at which point the gravitational wave frequency is in LISA's most sensitive range and is above the expected white dwarf background. Figure 9 shows the distribution of eccentricities for binaries with gravitational radiation in the LISA band. There are more low eccentricities at higher mass ratios. This is because at low mass ratios each encounter takes a fractionally larger amount of energy away from the binary than at high mass ratios. Thus at low mass ratios, the last encounter will tend to harden the binary such that it is closer to merger. At high mass ratios, however, encounters take a smaller fractional amount of energy from the binary, and, thus, the high mass ratio binaries have more time to circularize more during their orbital decay. For the 1000:10:10 mass ratio, a large fraction of the eccentricities are in the range 0.1 . e . 0.2 where the binary is eccentric enough to display general relativistic effects such as pericenter precession, but circular templates may be sufficient for initial detection of the gravitational wave. Finally, because the first few hundred million years of a cluster's life witness a large number of mergers, recently formed and nearby super star clusters are promising sources of gravitational waves from IMBH coalescence. -- 21 -- Fig. 9. -- Distribution of eccentricities after integrating the Peters (1964) equations until in LISA band when the orbital frequency νorb = 10−3 Hz. The solid histograms are the Newtonian only sequences, and the hatched histograms are sequences with gravitational radiation. The sequences with gravitational radiation tend towards lower eccentricity since they have already started to circularize during the sequence. There is more difference between the two cases at higher mass ratios since gravitational radiation is stronger. Higher mass ratio binaries have lower eccentricities than lower mass ratio binaries since the latter start closer to merger after the final encounter. The 1000:10:10 mass ratio shows that a large number of detectable binaries would have 0.1 . e . 0.2 such that they would likely be detectable by LISA with circular templates yet display measurable pericenter precession. -- 22 -- 6. Conclusions We present results of numerical simulations of sequences of binary-single black hole scattering events in a dense stellar environment. We simulate three-body encounters until the binary will merge due to gravitational radiation before the next encounter. In half of our simulations, we include the effect of gravitational radiation between encounters. 1. Sequences of high mass ratio encounters. Our simulations cover a range of mass ratios including those corresponding to IMBHs in stellar clusters. Because the binaries simulated are tightly bound, the encounters steadily shrink the binary's semimajor axis until it merges. The eccentricity, however, jumps chaotically between high and low values over the course of a sequence. Merger usually occurs at high eccentricity since gravitational radiation is much stronger then. 2. Gravitational wave emission between encounters. The inclusion of gravitational radiation between encounters affects the simulations in several ways. The extra source of shrinking caused by gravitational wave emission has the effect of shortening the sequence in terms of both the number of encounters and the total time, and the circularization from gravitational waves has the effect of decreasing the final eccentricity of the binary before it merges. 3. IMBH formation. Our simulations directly test the IMBH formation model of Miller & Hamilton (2002a). We find that there is sufficient time to build up to 1000 M⊙ when starting from 50 M⊙, but our simulations also show that if there are a thousand 10 M⊙ black holes in the globular cluster, the seed black hole would only be able to grow to 240 M⊙ before exhausting half of the black holes in the cluster. In addition, the probability of the binary's remaining in the cluster during a growth from 50 to 240 M⊙ is small. In order to avoid ejection from the cluster with a reasonable probability, either the black hole must have a larger mass at the onset of dynamical encounters, the cluster's escape velocity must be larger, or the black hole must grow by some additional mechanisms such as by Kozai- resonance-induced mergers, two-body captures, or from smaller interlopers. 4. Gravitational wave detection. The mergers of binary black hole systems are strong sources of detectable gravitational waves. We find that the merging binary will typically start with very high eccentricity. By the time the binary is detectable by the Advanced LIGO detector, it will have completely circularized, but when detectable by LISA, it may have moderate eccentricity (0.1 . e . 0.2) such that it will display general relativistic effects such as pericenter precession and still possibly be detectable with circular templates. We find a high rate of mergers in the first few hundred million years of a globular cluster. This suggests that recently formed, nearby super star clusters are promising sources for -- 23 -- gravitational radiation from IMBH coalescence. Further work in this study will be to include a distribution of interloper masses instead of a single population of 10 M⊙ black holes. A mass distribution of black holes is a more realistic model of a cluster core and could change the outcomes of the sequences. Exchanges will be more important since encounters with the more prevalent smaller black holes may do most of the hardening until a more massive black hole exchanges into the binary. We thank J. M. Fregeau, F. A. Rasio, and S. Sigurdsson for helpful discussions and comments. We are also grateful for the hospitality of the Center for Gravitational Wave Physics in which many fruitful ideas were born. Many of the results in this paper were obtained using the Beowulf cluster of the University of Maryland department of astronomy. This work was supported in part by NASA grant NAG 5-13229. REFERENCES Ando, M. et al. 2002, Classical Quantum Gravity, 19, 1409 Angelini, L., Loewenstein, M., & Mushotzky, R. F. 2001, ApJ, 557, L35 Barish, B. C. 2000, Adv. Space Res., 25, 1165 Baumgardt, H., Makino, J., Hut, P., McMillan, S., & Portegies Zwart, S. 2003, ApJ, 589, L25 Colbert, E. J. M., & Mushotzky, R. F. 1999, ApJ, 519, 89 Danzmann, K. 2000, Adv. Space Res., 25, 1129 Fabbiano, G., Schweizer, F., & Mackie, G. 1997, ApJ, 478, 542 Fabbiano, G., Zezas, A. L., & Murray, S. S. 2001, ApJ, 554, 1035 Favata, M., Hughes, S. A., & Holz, D. E. 2004 ApJ, submitted (astro-ph/0402056) Fidecaro, F., et al. 1997, in Proc. 12th Italian Conf on General Relativity and Gravitational Physics, ed. M. Bassan, V. Ferrari, M. Francaviglia, F. Fucito, & I. Modena (River Edge: World Scientific), 163 Fregeau, J. M., Joshi, K. J., Portegies Zwart, S. F., Rasio, F. A. 2002, ApJ, 570, 171 Fullerton, L. W., & Hills, J. G. 1982, AJ, 87, 175 -- 24 -- Gebhardt, K., et al. 2000, ApJ, 543, L5 Gebhardt, K., Rich, R. M., & Ho, L. C. ApJ, 578, L41 Gerssen, J., van der Marel, R. P., Gebhardt, K., Guhathakurta, P., Peterson, R. C., & Pryor, C. 2002, AJ, 124, 3270 Gurkan, M. A., Freitag, M., & Rasio, F. A. 2004, ApJ, in press (astro-ph/0308449) Heggie, D. C. 1975, MNRAS, 173, 729 Heggie, D. C., Hut, P., & McMillan, S. L. W. 1996, ApJ, 467, 359 Heggie, D. C., & Rasio, F. A. 1996, MNRAS, 282, 1064 Hut, P., & Bahcall, J. N. 1983, ApJ, 268, 319 Kaaret, P., Prestwich, A. H., Zezas, A., Murray, S. S., Kim, D.-W., Kilgard, R. E., Schelgel, E. M., & Ward, M. J. 2001, MNRAS, 321, L29 Kozai, Y. 1962, AJ, 67, 591 Madau, P., Rees, M. J. 2001, ApJ, 551, L27 Matsumoto, H., Tsuru, T. G., Koyama, K., Awaki, H., Canizares, C. R., Kawai, N., Mat- sushita, S., & Kawabe, R. 2001, ApJ, 547, L25 Meylan, G., Sarajedini, A., Jablonka, P., Djorgovski, S. G., Bridges, T., & Rich, R. M. 2001 AJ, 122, 830 Miller, M. C., & Hamilton, D. P. 2002, MNRAS, 330, 232 Miller, M. C., & Hamilton, D. P. 2002, ApJ, 576, 894 Peters, P. C. 1964, Phys. Rev. B, 136, 1224 Portegies Zwart, S. F., & McMillan, S. L. W. 2000, ApJ, 528, L17 Portegies Zwart, S. F., & McMillan, S. L. W. 2002, ApJ, 576, 899 Quinlan, G. 1996, New Astronomy, 1, 35 Schilling, R. 1998, in AIP Conf. Proc. 456, Laser Interferometer Space Antenna, Second International LISA Symp., ed. W. M. Folkner (New York:AIP), 217 Schneider, R., Ferrara, A., Natarajan, P., & Omukai, K. 2002, ApJ, 571, 30 -- 25 -- Sigurdsson, S., & Hernquist, L. 1993, Nature, 364, 423 Sigurdsson, S., & Phinney, E. S. 1993, ApJ, 415, 631 Sigurdsson, S., & Phinney, E. S. 1995, ApJS, 99, 609 van der Marel, R. P. 2003 Carnegie Observatories Astrophysics Series, Vol. 1: Coevolution of Black Holes and Galaxies, ed. L. C. Ho (Cambridge: Cambridge Univ. Press), in press van der Marel, R. P., Gerssen, J., Guhathakurta, P., Peterson, R. C., & Gebhardt, K. 2002, AJ, 124, 3255 Webbink, R. F. 1985, in Dynamics of Star Clusters, IAU Symposium 113, ed. J. Goodman & P. Hut (Dordrecht:Reidel), 541 This preprint was prepared with the AAS LATEX macros v5.2. -- 26 -- Table 1. Single Encounter Cross-sections for Exchange m0:m1:m2 Ejected Mass HHM96 This Work 10 : 1 : 1 10 : 1 : 10 3 : 1 : 1 1 10 1 10 1 3 1.054 ± .105 1.086 ± .023 -- -- -- -- 7.825 ± .360 0.520 ± .087 2.311 ± .170 0.059 ± .025 7.741 ± .255 0.513 ± .043 2.465 ± .073 0.072 ± .007 Note. -- This table compares dimensionless cross-sections for exchange ¯σ (see text for details) calculated by Heggie et al. 1996 and by us. The first column lists the masses, with binary components m0 and m1. Column two shows the mass of the ejected object. The ejection of the smaller mass is energetically favored so it always has a larger cross-section. There is general agreement between the two calculations to within the statistical uncertainty, which we calculate as ¯σ/N 1/2 ex , where Nex is the total number of exchanges. -- 27 -- Table 2. Sequence Statistics m0 Case hnenci hneji hfbineji htseqi /106 yr haf i /AU hef i 10 Newt. GR Evol. 20 Newt. GR Evol. 30 Newt. GR Evol. 50 Newt. GR Evol. 100 Newt. GR Evol. 200 Newt. GR Evol. 300 Newt. GR Evol. 500 Newt. GR Evol. 1000 Newt. GR Evol. 51.6 48.7 51.3 47.1 58.9 55.1 73.2 66.7 102.0 93.4 158.4 140.3 208.5 184.0 308.7 269.1 562.4 483.0 3.9 3.7 6.5 6.1 9.3 8.6 14.6 13.0 24.0 20.1 38.2 31.5 49.1 39.4 71.1 54.9 117.3 88.9 0.880 0.839 0.835 0.776 0.753 0.676 0.581 0.455 0.229 0.161 0.043 0.026 0.013 0.006 0.001 0 0 0 82.72 59.89 65.94 43.46 49.11 31.89 33.75 22.73 21.35 14.97 15.13 9.998 11.89 7.822 9.920 6.225 7.363 4.427 0.164 0.929 0.190 0.901 0.178 0.924 0.230 0.898 0.198 0.926 0.222 0.892 0.230 0.919 0.285 0.892 0.327 0.936 0.357 0.873 0.387 0.938 0.444 0.872 0.468 0.943 0.445 0.874 0.528 0.944 0.488 0.860 0.641 0.953 0.556 0.851 Note. -- Table 2 summarizes the main results of our simulations of sequences of three- body encounters. For each dominant mass, m0, we ran 1000 sequences of pure Newtonian encounters (Newt.) and 1000 sequences of the more realistic Newtonian encounters with gravitational radiation between encounters (GR Evol.). The columns list the average number of encounters per sequence hnenci, the average number of black holes ejected from the cluster in each sequence hneji, the fraction of sequences in which the binary is ejected from the cluster, hfbineji, the average total time for the sequence htseqi, the average final semimajor axis haf i, and the average final eccentricity hef i. -- 28 -- Table 3. IMBH Formation Seed Mass vesc Probability to remain (M⊙) (km s−1) in cluster Number of BH ejections Time (108 yr) 50.0 100.0 200.0 300.0 40.0 50.0 60.0 70.0 40.0 50.0 60.0 70.0 40.0 50.0 60.0 70.0 40.0 50.0 60.0 70.0 0.00264 0.0356 0.129 0.269 0.0821 0.290 0.525 0.698 0.670 0.842 0.932 0.978 1.000 1.000 1.000 1.000 7.06 6.15 4.93 4.05 6414 5276 4038 3573 6312 5188 3963 3606 5995 4922 4077 3417 5561 4564 3777 3164 Note. -- This table lists values for selected seed masses and cluster escape velocities for the gravitational radiation case. Column 3 lists the probability for the IMBH to remain in the cluster until it reaches a mass of 300 M⊙. The fourth column lists the total number of black holes ejected in building up to 1000 M⊙. Column 5 lists the total time to build up to 1000 M⊙. The total time is not affected by the escape velocity because the density of black holes in the cluster core is taken to be constant.
astro-ph/0305521
1
0305
2003-05-27T23:44:59
The Tully-Fisher Relation of Barred Galaxies
[ "astro-ph" ]
We present new data exploring the scaling relations, such as the Tully-Fisher relation (TFR), of bright barred and unbarred galaxies. A primary motivation for this study is to establish whether barredness correlates with, and is a consequence of, virial properties of galaxies. Various lines of evidence suggest that dark matter is dominant in disks of bright unbarred galaxies at 2.2 disk scale lengths, the point of peak rotation for a pure exponential disk. We test the hypothesis that the TF plane of barred high surface brightness galaxies is offset from the mean TFR of unbarred galaxies, as might be expected if barred galaxies are ``maximal'' in their inner parts. We use existing and new TF data to search for basic structural differences between barred and unbarred galaxies. Our new data consist of 2-dimensional Halpha velocity fields derived from SparsePak integral field spectroscopy (IFS) and V,I-band CCD images collected at the WIYN Observatory for 14 strongly barred galaxies. We use WIYN/SparsePak (2-D) velocity fields to show that long-slit (1-D) spectra yield reliable circular speed measurements at or beyond 2.2 disk scale lengths, far from any influence of the bar. This enables us to consider line width measurements from extensive TF surveys which include barred and nonbarred disks and derive detailed scaling relation comparisons. We find that for a given luminosity, barred and unbarred galaxies have comparable structural and dynamical parameters, such as peak velocities, scale lengths, or colors. In particular, the location of a galaxy in the TF plane is independent of barredness. In a global dynamical sense, barred and unbarred galaxies behave similarly and are likely to have, on average, comparable fractions of luminous and dark matter at a given radius. (abridged)
astro-ph
astro-ph
For publication in ApJ (v594). The Tully-Fisher Relation of Barred Galaxies St´ephane Courteau1 Department of Physics & Astronomy, University of British Columbia, 6224 Agricultural Road, Vancouver, BC V6T 1Z1 [email protected] David R. Andersen1 Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany [email protected] Matthew A. Bershady Univ. of Wisconsin, Dept. of Astronomy, 475 N. Charter St., Madison, WI 53706 [email protected] Lauren A. MacArthur1 Department of Physics & Astronomy, University of British Columbia, 6224 Agricultural Road, Vancouver, BC V6T 1Z1 [email protected] and Hans-Walter Rix Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany [email protected] 1Visiting Astronomer, Kitt Peak National Observatory, National Optical Astronomy Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc. (AURA) under cooperative agreement with the National Science Foundation. 1 -- 2 -- ABSTRACT We present new data exploring the scaling relations, such as the Tully-Fisher relation (TFR), of bright barred and unbarred galaxies. A primary motivation for this study is to establish whether barredness correlates with, and is a con- sequence of, virial properties of galaxies. Various lines of evidence suggest that dark matter is dominant in disks of bright unbarred galaxies at 2.2 disk scale lengths, the point of peak rotation for a pure exponential disk. We test the hy- pothesis that the TF plane of barred high surface brightness galaxies is offset from the mean TFR of unbarred galaxies, as might be expected if barred galaxies are "maximal" in their inner parts. We use existing and new TF data to search for basic structural differences between barred and unbarred galaxies. Our new data consist of 2-dimensional Hα velocity fields derived from SparsePak integral field spectroscopy (IFS) and V,I-band CCD images collected at the WIYN Observa- tory1 for 14 strongly barred galaxies. Differences may exist between kinematic and photometric inclination angles of barred versus unbarred galaxies. These findings lead us to restrict our analysis to barred galaxies with i > 50◦. We use WIYN/SparsePak (2-D) velocity fields to show that long-slit (1-D) spectra yield reliable circular speed measurements at or beyond 2.2 disk scale lengths, far from any influence of the bar. This enables us to consider line width measurements from extensive Tully-Fisher surveys which include barred and nonbarred disks and derive detailed scaling relation comparisons. We find that for a given luminosity, barred and unbarred galaxies have compa- rable structural and dynamical parameters, such as peak velocities, scale lengths, or colors. In particular, the location of a galaxy in the TF plane is independent of barredness. In a global dynamical sense, barred and unbarred galaxies behave similarly and are likely to have, on average, comparable fractions of luminous and dark matter at a given radius. Subject headings: galaxies: bars -- galaxies: -- galaxies: photometry -- galaxies: spirals -- galaxies: structure formation -- galaxies: kinematics 1The WIYN Observatory is a joint facility of the University of Wisconsin-Madison, Indiana University and the National Optical Astronomy Observatory. -- 3 -- 1. Introduction Based on the flatness of rotation curves in spiral galaxies and the density profiles inferred from X-ray temperatures and stellar velocity dispersion profiles in ellipticals, it is widely believed that galaxies are embedded in non-dissipative massive dark halos. More than 90% of the total mass of a galaxy would be in the form of dark matter. Less appreciated is the fact that we still have a rather muddled picture of the mass distribution of luminous and dark matter in the luminous part of a galaxy. This is unfortunate since the final distribution of baryons in a galaxy is a tell-tale sign of its formation and evolution. Numerical and analytical models of disk formation in a dissipationless dark matter halo predict, for realistic total fractions of baryonic to dark matter, that spiral disks should live in dark halos that dominate the mass fraction at nearly all radii (e.g. Mo, Mao, & White 1998), beyond about a disk scale length. This ratio may quite possibly be different for barred and unbarred galaxies of a given total mass or luminosity (Courteau & Rix 1999; hereafter CR99). Recent debates about the Cold Dark Matter paradigm (e.g. Weinberg & Katz 2002; Sellwood 2003; Courteau et al. 2003a) and galaxy structural properties inferred from new in- frared surveys (e.g. Eskridge et al 2002; MacArthur et al. 2003) have brought barred galaxies to the fore. Bar perturbations in galaxies, far from just being dynamical curiosities, actu- ally play a fundamental role in shaping galaxies into the structures we see today (see Buta, Crocker, & Elmegreen 1996 for reviews). For instance, the early dynamical evolution of a massive rapidly rotating gaseous bar could provide enough energy and angular momentum to significantly modify the inner CDM halo (Silk 2002). Dynamical and structural studies of barred galaxies are however few, due in part to the complexity in interpreting their ve- locity fields (e.g. Weiner et al. 2001) and their surface brightness profiles (e.g. MacArthur et al. 2003). Many large-scale flow studies of spiral galaxies have also excluded disturbed or barred galaxies to minimize scatter, as previously believed, in the distance-measuring tech- nique. The latter studies have enabled extensive scaling relation studies of unbarred galaxies, but little attention has been paid to their barred cousins. This is again deplorable as a com- parative study of the scaling relations for barred and unbarred galaxies would potentially unravel clues about the structure and origin of bars and the role of dynamical processes in establishing the Hubble sequence of disk galaxies. The body of numerical simulations of barred galaxies is comparatively richer and has recently reached new heights with the availability of superior N-body realizations with more than 106 particles (post 2010 readers may enjoy a moment of laughter). Until just recently, it was believed that bar instabilities in a disk might be suppressed by a massive halo. Thus only low concentration halos, or equivalently systems of very high surface brightness (HSB) or low angular momentum per unit luminosity, would be prone to generating a non-axisymmetric -- 4 -- (bar/oval) structure in their center (Ostriker & Peebles 1973; accounts of the misconceptions surrounding this argument are presented in Bosma 1996 and Sellwood & Evans 2001). The suggestion that barred galaxies would have an especially high ratio of baryons-to- dark matter within the optical disk ("maximal disk") might imply that these systems define their own sequence in the luminosity-line width diagram, if one assumes that unbarred galax- ies are, on average, sub-maximal (CR99). Thus, for a given absolute magnitude, a galaxy with higher baryon fraction, or disk mass-to-light (M/L) ratio, would have a shorter disk scale length and rotate faster. Verification of this important, though tentative, suggestion should be easily obtained from a large sample of uniformly selected barred galaxies that are part of a well-calibrated, self-consistent luminosity-line width survey. The current study was largely motivated by this question. In discussing the mass distribution in spiral galaxies, we shall use the definition that a disk is "maximal" if it contributes more than 75% of the total rotational support of the galaxy at Rdisk ≡ 2.2hdisk, the radius of maximum disk circular speed (Sackett 1997). Thus, for a maximal disk, Vdisk/Vtotal ∼> 0.75, where Vtotal is the total amplitude of the rotation curve at Rdisk and Vdisk = V (Rdisk). Note that for Vdisk/Vtotal = 0.7 the disk and halo contribute equally to the potential at Rdisk. Large bulges for late-type galaxies make little difference for the computation of this quantity at Rdisk (CR99). The pattern speeds of bars have been considered as a potential indicator of the relative fraction of dark matter in galaxy disks. N-body simulations of bar formation in stellar disks suggest that dynamical friction from a dense dark matter halo dramatically slows the rotation rate of bars in a few orbital periods (Debattista & Sellwood 1998, 2000; hereafter DS00). Because bars are observed to rotate quickly, DS00 proposed that dark matter halos in HSB galaxies must have a low central density; thus, their disks ought to be maximal. These simulations were revisited by Valenzuela & Klypin (2003; hereafter VK03) with similar N-body simulations (no gas) but with an order of magnitude improvement in the force resolution. VK03 found that dynamical friction from transfer of angular momentum of the bar to the halo does play a role but, contrary to DS00, that effect appears to be small. In addition, VK03 find that bars can form even in the presence of strong halos, and that stellar disks make a negligible contribution to the inner rotation curve (at Rdisk). The bars modeled in DS00 also span nearly the entire disk whereas the observed bar-to-disk scale length ratio seldom exceeds 1.5, as also pointed out by VK03. These authors find that mass and force resolution are critical for modeling the dynamics of bars, and the contentious results from DS00 would stem primarily from numerical resolution effects. However, the higher force resolution of VK03 induces numerical viscosity which may bring their results into question (J. Sellwood 2003; priv. comm.)! Free from the vagaries of numerical simulations, Athanassoula -- 5 -- (2003) uses analytical calculations to warn against the use of bar slowdown rate to set limits on the baryonic to dark matter fraction within the optical radius4, in agreement with Sellwood (2003). This point is however moot since one cannot observe bar slowdown from a single snapshot. Athanassoula (2003) further suggests that the ratio of corotation to bar length may not be an adequate estimator of halo fraction, but the model ratios (see her Figs. 12 & 14) upon which these conclusions are drawn are inconsistent with observations. A complete picture of bar dynamics awaits a self-consistent treatment of both the stars and gas embedded in a cosmologically motivated halo. These simulations should include dynamical friction and ultimately reproduce the fraction of strong bars detected in the infrared and predict the rate of bar slowdown and dissolution as a function of bulge/total brightnesses, time, and environment. The model-independent quest of the relative matter distribution in barred and unbarred galaxies is by no means straightforward either, but is most significant as it provides a nec- essary constraint for the shape and amplitude of the dark matter density profile in the luminous part of a galaxy. Whether disks are maximal or not at Rdisk, the inner 1-2 kpc may be dominated by baryons in most galactic systems, including early and late-type HSB barred and unbarred spirals (e.g. Broeils & Courteau 1997; Corsini et al. 1999), low surface brightness (LSB) galaxies (Swaters 1999; Swaters, Madore, & Trewhella 2000; Fuchs 2002) and ellipticals (e.g. Brighenti & Mathews 1997; see also Ciotti 2000). Maximally massive disks in LSB galaxies may however require unrealistically high disk M/L ratios (Swaters et al. 2000; Fuchs 2002), based on stellar population synthesis models. Also troublesome is our lack of knowledge about the distribution of matter in our own Milky Way. Whether it has a maximal disk (Gerhard 2002) or not (Dehnen & Binney 1998; Klypin, Zhao, & Somerville 2002) is still a matter of debate. Crucial elements for local mass density estimates include the precise contribution of the massive central bar (e.g. Zhao, Rich, & Spergel 1996) or elongated bulge (Kuijken 1995), an accurate measure of the disk scale length, and constraints from microlensing towards the bulge. The determination of the relative fraction of visible and dark matter in external barred and unbarred galaxies relies on our ability to determine stellar M/L ratios accurately. The modeling of disk dynamical mass in barred galaxies relies heavily on the interpretation of the non-axisymmetric motions of ionized gas around the bar within the context of a hydrodynamical model. This model does have a local potential, and hence the bar and disk M/L are parameters of the model. It is certainly a more complicated approach than 4Athanassoula (2003) finds that the bar slowdown rate depends not only on the relative halo mass at a given radius, but also on the velocity dispersion of both the bulge and disk components. -- 6 -- using collisionless particles as dynamical tracers, as with stellar velocity dispersions, but the latter has its own complications as well (e.g. Swaters et al. 2003). Significant improvements in mass modeling techniques for individual galaxies are expected with the development of stellar population synthesis models (Bell & de Jong 2001) and dynamical constraints (Weiner, Sellwood, & Williams 2001) to yield realistic M/L ratios, and further constraints from cosmological simulations of dark halos to curtail disk-halo degeneracies (Dutton, Courteau, & de Jong 2003). Various lines of circumstantial evidence for external systems favor dark matter halos that dominate the mass budget within Rdisk. Arguments based on the stellar kinematics of galactic disks (Bottema 1997), gas kinematics (Kranz, Slyz, & Rix 2003), the stability of disks (Fuchs 2001) and the lack of correlated scatter in the Tully-Fisher relation (hereafter TFR; Tully & Fisher 1977) of unbarred LSB and HSB galaxies (CR99) suggest that, on average, disks with Vmax < 200 km s−1 are sub-maximal. The two very different analyses by Bottema and CR99 both yield Vdisk/Vtotal = 0.6 ± 0.1, or Mdark/Mtotal = 0.6 ± 0.1 for HSB galaxies at Rdisk. The geometry of gravitational lens systems, coupled with rotation curve measurements, can also be used to decompose the mass distribution of a lensing galaxy. This promising technique, pioneered by Maller et al. (2000), has been applied to the galaxy-lens system 2237+0305 by Trott & Webster (2002) who find Vdisk/Vtotal = 0.57±0.03, in excellent agreement with the studies above and predictions from analytical models of galaxy formation (e.g. Dalcanton et al. 1997; Mo, Mao, & White 1998). While a consistent picture of galaxy structure is emerging in which a dark halo dominates with Mdark/Mtotal ≥ 0.6 well into the optical disk, a number of pro-maximal disk arguments are still found in the literature, citing evidence from the shapes and extent of rotation curves and mass modeling (see, e.g., Bosma 2002). The match between pure disk mass models and Hα rotation curves (e.g. Broeils & Courteau 1997; Seljak 2002; Jimenez et al. 2003) is usually satisfactory for spiral galaxies of different surface brightnesses and morphologies and has often been invoked as evidence for a maximal concentration of baryons relative to the dark matter inside the optical disk (Buchhorn 1992; Palunas & Williams 2000). However, mass modeling with Hα rotation curves alone is not a uniquely determined problem. The equivalence of, or degeneracy between, the two descriptions -- pure disk versus sub-maximal disk + dark halo -- was demonstrated in Broeils & Courteau (1997) and CR99 for a sample of 300 disk galaxies; residuals for the maximal or sub-maximal fits are indistinguishable. Without an accurate estimate of M/Ldisk, or external constraints on Vdisk/Vvirial at Rdisk, mass modeling cannot dissentangle maximal and sub-maximal disk models. Our study of the dynamical structure of barred and unbarred galaxies will offer new insights in the debate of the maximal disk hypothesis in barred and unbarred galaxies. -- 7 -- However, we plan to revisit this controversial issue in a future presentation (Courteau et al. 2003a). Here we pursue our comparison of barred and unbarred galaxies in the context of global scaling relations. 1.1. Available galaxy samples The study of scaling relations of barred galaxies, and tracing their location in the TFR, requires that we utilize "fundamental plane" surveys of an ensemble of galaxies. The "Shellflow" and "SCII" all-sky Tully-Fisher surveys of Courteau et al. (2000) and Dale et al. (1999) are useful in that respect. These surveys were designed to map the convergence of the velocity field on ∼ 60h−1 Mpc scales while minimizing calibration errors between dif- ferent telescopes in different hemispheres; state-of-the-art TF calibrations are thus available in both cases. Both surveys include line width and luminosity measurements for a small fraction of barred galaxies that can be used to study structural trends, provided the pres- ence of bar does not bias these measurements. More details about the surveys will be given in §3. In order to calibrate existing long-slit spectra of barred galaxies and initiate a com- prehensive study of barred galaxy velocity fields, we have collected new deep V and I-band images and integral field Hα velocity fields of 14 strongly barred galaxies at the WIYN 3.5-m telescope. We present the new data and velocity field analysis in §2 and discuss possible limitations of the data, such as due to inclination uncertainties and non-circular motions. We then examine the location of barred and unbarred galaxies in the TF samples discussed above in §3. We find that barredness does not play a role in the luminosity-line width and luminosity-size planes of spiral galaxies. In §4, we discuss future programs that may benefit the study of scaling relations in barred and unbarred galaxies. 2. A New WIYN Survey of Barred Galaxies 2.1. Observations In March 2002, we obtained 2-D Hα velocity maps and deep V and I-band photometry at the WIYN 3.5-m (3 nights) and WIYN 0.9-m (2 nights), respectively for 14 strongly <∼ 15; see Table 1) and one unbarred spiral galaxy barred bright galaxies (SBb-SBc; mB (NGC 3029). The galaxies were selected according to the same criteria as the TF Shellflow survey of spiral galaxies, save the emphasis on the bar-like morphology. Ultimately, we aim to calibrate our new data on the same system as Shellflow, a survey deficient in barred -- 8 -- galaxies, to enable direct comparisons between barred and unbarred systems. Integral field spectroscopy (IFS), which is lacking in Shellflow and SCII, is required to fully characterize the velocity amplitudes of the bulge, bar, and underlying disk, especially if non-circular velocities are conspicuous. We have obtained 2-D velocity maps with the SparsePak integral field unit (Bershady et al. 2003a). The SparsePak IFU is a fiber-optic array of 82 fibers mounted at the Nasmyth f/6.3 focus imaging port on the WIYN 3.5-m telescope. SparsePak has 75 fibers arranged in a sparsely filled grid sub-tending an area of 72′′× 73′′. Each fiber has an active core diameter of 4.′′69 (500 µm); cladding and buffer increase the total fiber diameter to 5.′′6. The filling factor for the grid is ∼ 25% on average, but rises to ∼ 55% in the inner 16′′ where the fibers are more densely packed. In addition to the 75 fibers arranged in a square, another 7 fibers are spaced around the square roughly 70′′ -- 90′′ from the center and are used to measure the "sky" flux. An example of the SparsePak footprint is shown in Fig. 1. SparsePak feeds the WIYN Bench Spectrograph, a fiber -- fed spectrograph designed to provide low to medium resolution spectra. We used the Bench Spectrograph camera (BSC) and 316 lines mm−1 echelle grating in order 8 to cover 6500A < λ < 6900A, with a dispersion of 0.2A pix−1 (8.8 km s−1 pix−1) and an instrumental FWHM of 0.6A (26.5 km s−1). The BSC images the spectrograph onto a T2KC thinned SITe 2048 × 2048 CCD with 24 µm-pixels. The chip has a read noise of 4.3 e− and was used with the standard gain of 1.7 e−/ADU. The peak system throughput for this setup is roughly 5.5%, estimated from standard-star observations (Bershady et al. 2003b). Given SparsePak's ∼ 15′′ center-to-center fiber spacing and total area, we used 3 point- ings along the galaxy's position angle to maximize spatial coverage and filling factor. Typical pointing offsets were ∼ 6′′. The observed galaxies have moderate sizes (a ∼ 2.′0) and their velocity field can thus be mapped from center-to-edge. Total Sparsepak integrations con- sisted of 3 pointings × 2 900s exposures per pointing, for a total of 1.75 hours per galaxy. Multiple exposures at each position were used to identify and remove cosmic rays. Spectra obtained from SparsePak closely resemble WIYN Densepak or Hydra spectra (i.e. multi-fiber spectral data). Thus basic spectral extraction, flattening, wavelength cali- bration and sky subtraction were done using the NOAO IRAF5 package dohydra. After basic reductions, we used a Gaussian line -- fitting algorithm to measure Gaussian fluxes, widths, centers and centroid errors for Hα emission lines (Andersen et al. 2003). We rejected any 5IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the As- sociation of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. -- 9 -- line with a S/N < 5. More than 70% of measured Hα lines, even at the edge of the field, had significantly higher signal-to-noises, with S/N ∼> 20, yielding a mean centroiding error of only 2.4 km s−1 for these 15 galaxies. The V and I images were acquired at the WIYN 0.9-m telescope in f/13.5 mode Isophotal brightness errors are ∼< 0.1 (0.′′43 pix−1); integrations were 600s in each filter. mag arcsec−2 at, or below, 26.5 mag arcsec−2 in V and I. The imaging was obtained in non- photometric conditions (thin wisps covered the Arizona desert sky) and thus cannot readily be merged into the Shellflow imaging data base. Structural parameters can still be measured accurately, down to deep levels, as we discuss below. Three previously observed SB (NGC 2540, UGC 5141, UGC 8229) and two SAB (NGC 3029, UGC 6895) Shellflow galaxies with available long-slit Hα spectra and V,I photometry were duplicated at the WIYN telescopes for comparison. These observations enable us to tie the SparsePak velocity field information with the Shellflow long-slit spectra obtained with the KPNO & CTIO 4m telescopes + RC Spec (Courteau et al. 2003b). 2.2. Data Analysis Azimuthally-averaged surface brightness profiles were extracted for all the galaxies using ellipse fitting with a fixed center. To ensure a homogeneous computation of structural parameters and color gradients, we use the position angles and ellipticities of our I-band isophotal maps to determine the SB profiles in the V-band. The position angle and ellipticity are allowed to vary at each isophote. Please refer to Courteau (1996) for details about our surface brightness extraction technique. Reduction techniques for the extraction of rotation curves from long-slit spectra are de- scribed in Courteau (1997). We shall simply state that the 1-D rotation curve is constructed by measuring an intensity-weighted centroid at each resolved major-axis Hα emission feature above a noise threshold. For the 2-D SparsePak data, a single, inclined, differentially rotat- ing, circular disk model with a fixed center is used to fit the Hα velocity fields (Andersen & Bershady 2003). Briefly, we assume a radially symmetric rotation curve and an axisymmetric velocity field. Using this smooth functional representation of the velocity field, we compared the model velocity field to observations. Parameters are varied using a multi-dimensional down-hill simplex method (Press et al. 1992) to minimize a χ2 statistic. Our velocity field model has nine free parameters: seven for the rotation curve (see next paragraph), and two for inclination and position angle. Two additional parameters account for positional offsets from differential telescope pointing errors for each SparsePak position, yet in practice these -- 10 -- parameters were consistent with zero and were thereafter not allowed to vary. We parameterize the model used to fit the rotation curves of both the 1-D (long slit) and 2-D (SparsePak) velocity field data with the following empirical function (Courteau 1997): v(r) = v0 + va (1 + x)β (1 + xγ)1/γ , (1) where x = 1/R = rt/(r − r0), v0 and r0 are the velocity and spatial centers of rotation, va is an asymptotic velocity, and rt is a transition radius between the rising and flat part of the rotation curve. Solid-body rotation, or v(r) ∝ r (with ∂v/∂r ∼ va/rt), is recovered for r −r0 ≪ rt, and flat rotation, or v(r) ∝ va, is achieved for r −r0 ≫ rt. The term γ governs the degree of sharpness of turnover, and β can be used to model the drop-off or steady rise of the outer part of the rotation curve. Table 1 gives velocity field and structural parameters for the SparsePak data collected at WIYN in March 2002. Listed are the number N of velocity data points, the kinematic and photometric inclinations, the kinematic and photometric position angles, the velocity fit parameters, va, rt, β, and γ (see Eq. 1), the bar radius, Rbar in the plane of the galaxy, the I- band scale length h of the disk, and the recessional velocity of the galaxy, v0. The bar radius is defined as the location where the I-band surface brightness drops and/or position angle changes abruptly. Disk scale lengths were determined as in MacArthur et al. (2003). No photometric parameters are listed for IC 0784 which could not be observed at the telescope due to time and weather constraints. Appendix A contains rotation curves and extracted velocity fields (spider diagrams) for the WIYN/SparsePak galaxies. The model rotation curves, based on Eq. (1), are a decent match to most extracted integral field velocity data points. These models are shown mostly for illustrative purposes and for comparison with similar fits to rotation curves derived from long-slit spectra. They can also be used for future dynamical modeling. The overall impression from the comparison of velocity data for the 5 Shellflow galaxies with long-slit 1-D and SparsePak 2-D rotation curves in Appendix A is very favorable. For NGC 2540 (Fig. 9), the 1-D and 2-D velocity models are indistinguishable, owing in part to the very similar position angles and inclinations used to extract the velocity amplitudes. The unbarred galaxy, NGC 3029 (Fig. 10), was re-observed for consistency check; again the velocity data and models agree very well within the measurement uncertainty. NGC 5141 shows only slight differences in the modeled RCs, and UGC 6895 and UGC 8229 show slightly larger differences in the inner slopes, perhaps caused by a misaligned slit. While the data distributions agree within their respective scatter, the RC models predict different maximum rotation speeds, at the 10-20 km s−1 level. However, the basic impression to retain for this -- 11 -- comparison is that long-slit and IFS rotation curves agree well within their measurement errors and intrinsic scatter and it can be assumed that line widths from 1-D rotation curves are a fair representation of the overall velocity field, even for barred galaxies. Close agreement between 1-D rotation curves from Hα long-slit spectra and major-axis rotation curves from Fabry-Perot (2-D) velocity fields was also demonstrated by Courteau (1997). Another concern, when mapping the kinematic and dynamic structure of barred galax- ies, is whether our diagnostics are affected by non-circular velocities, radial flows, and/or isophotal distortions. In order to assess the importance of non-circular motions, we have examined minor-axis rotation curves (not shown here for simplicity) and spider diagrams in Appendix A (see also Swaters et al. 2003). The minor-axis rotation curves are consistent with 10 -- 20 km s−1 velocity dispersions of the turbulent gas with little hint of systematic deviations. The spider diagrams do show signs of non-circular motions, especially within ∼ 1.2Rbar(≃ 1.5hdisk). However, beyond the extent or reach of the bar, most position- velocity diagrams are symmetric about the major kinematic axis. With the exception of IC 2104 (Fig. 8), a symmetric velocity pattern is recovered for all galaxies at, and beyond, Rdisk. The good match between 1-D and 2-D velocity fields and lack of significant non-circular motions at or beyond Rdisk suggests that we can compare raw rotation speeds of barred and unbarred galaxies, all other quantities being equal, without significant bias. This is what we do in §3 for the Shellflow and SCII data. Any putative offset of the barred galaxies in the TF plane should not be due to systematic effects in the line widths. Deprojection of velocity fields requires an inclination estimate. TF studies usually make use of photometric inclinations determined in the outer disk, away from a bar or spiral dis- tortions, where ellipticities and position angles do not vary appreciably (e.g. Courteau 1996, Beauvais & Bothun 2001). We compare our SparsePak kinematic and I-band photometric inclination and position angle estimates in Fig. 2 and Table 1. A position angle offset would systematically lower the observed long-slit rotation, and inclination differences could displace a galaxy in the TF plane. We find that galaxies with ikin > 45◦ show no appreciable incli- nation offset (within 3◦ rms) and a mild position angle offset (10◦ rms) between kinematic and photometric estimates. Position angle differences can be large for more face-on galaxies but our sample is too small to isolate systematic trends. For galaxies with i < 35◦, photometric inclination angles are, on average, ∼ 12% larger (more edge-on) than kinematic estimates. Inclination offsets for the low-inclination unbarred galaxy NGC 3029 are large and can only be explained by model fitting (kinematic vs isopho- tal) differences, whereas excellent agreement is found for UGC 6895, a higher inclination (i = 45◦) unbarred galaxy. -- 12 -- Note that our velocity model assumes circular, instead of elliptical, orbits. Kinematic inclinations are still precise enough to construct a TFR with small scatter (σTF ≃ 0m. 3) even at very low inclinations (Andersen & Bershady 2003). It is however unclear which of the kinematic or photometric inclination is more "representative" of the disk projection on the sky. The inclination offset may result from a combination of kinematic modeling that favors more circular orbits and great sensitivity of the isophotal mapping technique to m = 2 brightness perturbations. Spiral arms typically originate at the ends (ILR) of bars and retain a small pitch angle, highly noticeable in the brightness distribution, hence the plausible bias towards higher photometric inclinations. These effects are especially acute when spiral arms are fully resolved. In a similar study, Sakai et al. (2000; H0 Key Project) find that photometric and kine- matic (radio synthesis mapping) inclination angles differ for barred galaxies. Among the 21 calibrator galaxies in their TF sample, 7 are barred and their kinematic inclination angles are ∼ 10-15% smaller than photometric inclinations. Their barred galaxies all have iphot > 45◦. However, inclination offsets for their unbarred galaxies are nearly absent. Peletier & Willner (1991) give radio and infrared inclination angles for 13 barred and unbarred nearby spirals with 27◦< i <70◦. Radio synthesis inclinations are also ∼12◦ smaller than photometric estimates, but for all inclinations. To illustrate this potentially confusing situation, we plot in Fig. 3 the inclination dif- ference, ∆i (kin -- phot), against kinematic inclination for the galaxy samples considered above, plus a sample of nearby, face-on, unbarred spiral galaxies (Andersen 2001). At low inclinations, kinematic inclinations appear to be systematically lower (more face-on) than photometric inclinations, with a trend of increasing differences with decreasing inclination. This is made very clear by examination of Andersen's data. At high inclinations, both barred and unbarred galaxies have smaller inclinations offsets, apparently independent of inclination. At these high inclinations, the effect on the velocity deprojection is negligible (< 5%). It may be that SparsePak and photometric inclinations in these inclined galaxies are affected by extinction as higher opacity would naturally bias high optically-determined inclinations. However, the radio synthesis inclinations compiled in Sakai et al. are insensitive to dust and the inclination difference is most likely explained by modeling differences; 2-D velocity fields are modeled under the assumption of circular orbits and the larger kinematic inclinations at large inclination may result from an underestimate of the disk thickness. In general, with increasing inclination, photometric inclinations become increasingly sensitive to the estimated disk thickness while velocity fields (especially radio velocity fields) become increasingly affected by warps and other non-circular motions. In any event, the inclination differences at ikin > 50◦ are small (< 5◦) and do not affect our study. Barnes & Sellwood (2003) find a similar result for a sample of inclined galaxies with inferred photometric and -- 13 -- kinematic (Fabry-Perot) inclinations. Opposite trends are found in the compilation of Peletier & Willner (1991) if all their data are considered. Inclination offsets are large even at high inclinations. This discrepancy however hinges on three galaxies, NGC 4178, 4192, and 4216, that display various patholo- gies. NGC 4178 is a very late type system, NGC 4192 has a strong warp in the outer disk, and NGC 4216 has a very pronounced dust-lane; these all make photometric measurements uncertain. If we ignore the Peletier & Willner data (bottom panel; Fig. 3), we find that the transition threshold where kinematic inclinations becomes significantly lower than pho- tometric inclinations depends on type: ikin = 50◦, 40◦, and 30◦ for barred, weakly-barred, and unbarred galaxies. Clearly, a more extensive 2-D spectroscopic survey of barred and unbarred galaxies in the near-infrared and radio will help address our general concerns about their dynamical structure and the limitations of our modeling techniques. Infrared imaging should also be secured for extinction-free inclination measurements. The measurement of a "true" inclina- tion of a galaxy is certainly ill-defined as it depends on the bandpass, dust extinction, the detector, reduction methods, and assumptions concerning the galaxy structure (e.g. pres- ence of warps). Yet, inclination angles from radio synthesis mapping may come closest to the most representative tilt angle of a galaxy on the sky. As we await more detailed comparisons of radio and optically determined inclinations, systematic differences between barred and unbarred galaxies can be avoided if we restrict our Shellflow and SCII samples to galaxies with iphot ∼> 50◦. Fortunately, all barred galaxies in our samples (Shellflow, SCII) already meet this criterion. We pursue our TF analysis with a discussion of Shellflow and SCII galaxies below. Our SparsePak sample will be reconsidered for TF analysis when calibrated imaging is available. 3. The Tully-Fisher Relation of Barred Galaxies We use the "Shellflow" and "SCII" all-sky TF surveys to map the location of barred galaxies in the TF plane. Shellflow includes 300 bright spiral (Sab-Scd) field galaxies in a shell bounded at 4500 < cz < 7000 km s−1, and SCII has 441 cluster spirals (Sa-Sd) spanning 5000 < cz < 19000 km s−1. Shellflow galaxies were drawn from the Optical Redshift Survey sample of Santiago et al. (1995) with inclinations in the range [45◦,78◦], mB ≤ 14.5, and b ≥ 20◦. Inter- acting, disturbed, and some barred galaxies were rejected. Rotation speeds from resolved -- 14 -- Hα rotation curves were measured at 2.2 disk scale lengths; the upper inclination limit (i < 78◦) reflects a desire to minimize extinction effects in the inner parts of the rotation curve (e.g. Courteau & Faber 1988; Giovanelli & Haynes 2002). Deep I-band and V-band images were collected for each Shellflow galaxy. Disk scale lengths were obtained from B/D decompositions of the azimuthally-averaged I-band surface brightness (SB) profile (Courteau et al. 2003b). The SCII cluster galaxies were selected from CCD I-band images taken at the KPNO and CTIO 0.9-m telescopes and classified by eye and by their bulge-to-disk ratio or concentration index. These galaxies have inclinations in the range [32◦,90◦] and I-band magnitudes 12 ≤ mI ≤ 17. SCII line widths were measured from both Hα long-slit spectra and HI line profiles. SCII disk scale lengths were obtained by "marking the disk", or fitting the exponential part of the SB profile from ∼21 I-mag arcsec−2 to ∼25 I-mag arcsec−2 (Dale et al. 1999). ≤ −24 and −18 ≤ M SCII I I Shellflow and SCII galaxies have −20 ≤ M Shell ≤ −24, re- spectively. Both TF calibrations are based on digital I-band imaging; V − I colors, to test for M/L variations and extinction effects, are available for the Shellflow sample only. De- projection of velocity widths uses photometric inclinations measured in the outer disk where ellipticities and position angles do not vary appreciably. Shellflow and SCII magnitudes are corrected for Galactic and internal extinction and distances account for a Hubble expansion, bulk flow model, and effects of incompleteness. The exact choice of distance scale does not affect our conclusions. According to the RC3, 37% of the Shellflow sample is barred (SB types only). In general, the proportion of galaxies with bars of all sizes is even higher (Eskridge et al. 2002) but we are here only concerned with galaxies with the strongest bars; i.e. those with potentially the highest central baryon fraction. Visual examination of the Shellflow galaxies revealed only 6 strongly barred systems (at I-band); these have Rbar/hdisk ≥ 1.2, where Rbar and hdisk are the size of the bar semi-major axis and disk scale length, respectively. Visual examination of the SCII galaxies yielded 27 strongly barred galaxies (D. Dale 2002; priv. comm.) In both samples, only barred galaxies with MI ≤ −20.4 could be identified. The Shellflow and SCII sub-samples of barred galaxies are by no means complete, nor are the parent catalogs, and a significant number of bars will be missed especially at low magnitudes and high inclinations where morphological identification becomes problematic. Figs. 4 & 5 show the distributions of rotational velocities and exponential scale lengths vs I-band absolute magnitudes for Shellflow and SCII galaxies. Different symbols identify the full range of spiral Hubble types; barred galaxies are further emphasized as solid symbols with open circles. Looking at the upper panel of Fig. 4 for Shellflow galaxies, one sees a small offset of barred galaxies from the mean TFR, consistent with these galaxies being -- 15 -- systematically brighter for a given mass (line width). The same statistically loose trend for barred galaxies was observed by Sakai et al. (2000). It could be explained if barred galaxies have higher star formation rates. However, Phillips (1996) and Kennicutt (1999) find that global star formation rates in barred and unbarred galaxies of the same Hubble type are comparable. The TF offset, if real, might also be consistent with maximal disks being brighter than their dark-matter dominated counter-parts at a given mass. A clearer picture is obtained with the larger SCII sample (Fig. 5) which shows no offset from the mean TFR for SCII barred galaxies. The combined velocity offset for the Shellflow and SCII barred galaxies in the two samples is hδlogV i = −0.02 ± 0.04, consistent with no deviation of the mean TFR. Note that photometric inclinations are used to deproject velocities in Shellflow and SCII but using kinematic inclinations instead would simply imply a readjustment of the TF zero-point. Provided only one inclination measure is used, the relative distribution of barred and unbarred TF galaxies is not affected by the precise choice of inclination (§2.2). Recall that all the Shellflow and SCII barred galaxies have i > 50◦ and are not affected by a putative (kin-phot) inclination offset. Furthermore, if we exclude the few unbarred galaxies that have i < 50◦ from the Shellflow and SCII samples, the TF distributions remain the same. Thus, we conclude that barred galaxies lie on the same TFR as unbarred galaxies. A similar realisation was also reached by Debattista & Sellwood (2000). The kinship between barred and unbarred galaxies extends to other properties as well. The lower panels of Figs. 4 & 5 show no statistical differences in the scale lengths of barred and unbarred galaxies (for a given absolute magnitude). Fig. 6 shows the color-magnitude diagram of Shellflow galaxies. Notwithstanding small statistics, barred and unbarred galaxies have similar colors, consistent with their having comparable star formation rates (Kennicutt 1999). MacArthur et al. (2003) find other similarities for structural parameters of barred and unbarred galaxies: their bar/bulge light profiles are close to exponential, and their ratio of bulge effective radius, re, and disk exponential scale length, h, falls in the range re/h = 0.22 ± 0.09, expected for late Freeman Type I spirals. CR99 developed and applied a test for correlated scatter of the TFR. According to this test, pure stellar exponential (maximal) disks should deviate from the mean TF and luminosity-size (LS) relations in such a way that ∂ log Vdisk / ∂ log Rexp = −0.5. Thus, strongly correlated TF/LS residuals for the barred spirals would support the suggestion that unbarred spirals have sub-maximal disks (high concentration halos) and that maximal disks are only found, on average, in barred spirals. A new analysis based on the Shellflow and SCII data sets yields residuals that are consistent with ∂ log Vdisk / ∂ log Rexp = 0.0 for both barred and unbarred galaxies. CR99 found a similar result for the Courteau-Faber sample. This result further confirms earlier observations about spiral galaxies: barred and -- 16 -- unbarred galaxies have similar physical properties and populate the same TF/LS relation and residual space. It also shows that the TFR is fully independent of surface brightness (CR99), a situation which may also result from the fine-tuning of virial parameters. The analysis of the independence of surface brightness in the TFR, and a revised interpretation of the "Courteau-Rix" test in terms of virial parameter correlations, is presented in Courteau et al. (2003a). 4. Discussion and Conclusion We have tested the hypothesis that barred and unbarred spiral disks have different structural correlations, such as the Tully-Fisher relation, with barred galaxies possibly hav- ing a higher luminous-to-dark matter fraction in their inner parts. New WIYN/SparsePak integral field spectroscopy and deep near-infrared photometry of barred and unbarred spirals allowed us to verify that non-circular motions are not significant at Rdisk and that rotation curves from 1-D or 2-D spectroscopy are reliable beyond that radius. Based on this result, and uniform inclination corrections for spiral galaxies with i > 50◦, we have compared the distribution of barred and unbarred galaxies in the TF plane from extensive redshift-distance surveys of galaxies and found no significant differences. For a given circular velocity, barred and unbarred galaxies have comparable luminosi- ties, scale lengths, colors, and star formation rates6. This suggests that barred and unbarred galaxies are close members of the same family and do not originate from different evolution- ary trees. Their structural duality may be understood if bars are generated by transient dynamical processes that are likely independent of the initial galaxy formation conditions. Their virial properties would otherwise be different. Very recent N-body simulations with the highest resolution have relaxed the notion that bars would grow in structures defined by a narrow range of disk/halo parameters. Thus our comparisons cannot be used to ascertain the notion that bars live mostly in spiral disks whose stellar fraction dominates the mass budget within the optical disk. Our results are however consistent with bright barred galaxies having similar dark matter fractions as do their unbarred cousins (Debattista & Sellwood 2000; Courteau et al. 2003a). Stellar velocity dispersions, which provide robust disk M/L ratios, hold the promise of breaking the disk/halo degeneracy in mass modeling of barred and unbarred galaxies. 6A comparative study Sheth et al. (2002) of the molecular gas properties of barred and unbarred galaxies in the BIMA Survey of Nearby Galaxies shows striking differences. However, their data (see their Fig.2) show less striking differences for the star formation rates between barred and unbarred galaxies but based on scanty information. More data are clearly needed to elucidate these questions! -- 17 -- If the presence of bars in rotating disks is not directly related to their virial structure but rather to their local dynamical state, it can surely be used as a signpost of galaxy evolution. Given that bars may be just as important as mergers in shaping field disk galaxies, significant efforts should be invested in programs to probe differences between barred and unbarred galaxies. Bars, which can be triggered spontaneously by the global dynamical instability of a rotationally supported disk, can also be induced by interactions with a satellite. One might thus expect an increase of the fraction of barred disks at higher redshift, unless these younger disks are too dynamically hot to sustain bar unstable modes. Van den Bergh et al. (2002) studied the visibility of bars in the northern Hubble Deep Field (HDF-N) and reported a dearth of bars at z > 0.7 in the rest-frame V-band. Taken at face value, this could indicate a dependence of bar strength on the local galaxy density which grows with time. However, a similar study by Sheth et al. (2003) based on the NICMOS Deep Field reveals numerous strongly barred galaxies up to z = 1.1. Extinction effects in the bluer band explored by van den Bergh et al. (2002) thus thwarted their ability to detect dust enshrouded bars. Given the detection of stable disks beyond z ∼ 1.3 (van Dokkum & Stanford 2001; Genzel et al. 2003), it is thus reasonable to posit the existence of bars at comparably high redshifts. The cosmological volumes sampled in two HDF studies above are very small and robust statistics on the barredness of galaxies with look-back time awaits wider coverage and more extensive sky surveys, especially with telescopes like ALMA within the next decade. Closer to home and on shorter time scales, our comparison of a few dozen barred galaxies with TF samples of unbarred disks should soon be superceded, it is wished, by systematic studies of structural and environmental properties of thousands of barred and unbarred galaxies in the SLOAN and 2MASS galaxy catalogs. Only with such large-scale, systematic local investigations can we make significant progress in mapping galaxy evolution at high- redshift and linking the near and far-field Universe. We are grateful to Daniel Dale for information about the SCII data base and to Anatoly Klypin, Jerry Sellwood and Ben Weiner for comments about N-body simulations of barred galaxies. Constructive suggestions by the referee improved the flow of the paper. This research has made use of the NASA/IPAC extragalactic database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. SC wishes to acknowledge his colleagues on the Shellflow team (Marc Postman, David Schlegel, and Michael Strauss) for permission to use previously unpublished results. SC and LAM acknowledge financial support from the National Science and Engineering Council of Canada. MAB acknowledges financial support from NSF grant AST-9970780. SC would also like to thank the Max-Planck Institut fur Astronomy in Heidelberg and the Max-Planck Institut fur Astrophysik in Munich for their hospitality while much of this paper was cooked up. -- 18 -- A. Rotation Curves, Velocity Fields and I-band images for the WIYN02 Sample This section shows long-slit (1-D) and SparsePak (2-D) velocity fields for all galaxies observed at WIYN in March 2002. See §2 for details about the sample and data analysis. Shown for each galaxy are the position-velocity contours ("spider" diagrams) in the upper window, superimposed on the galaxy I-band image, and in the lower window, the SparsePak velocities (in the plane of the sky, i.e. not corrected for projection effects). The velocity data were extracted according to various techniques described in the text and, whenever available, matching rotation curves are shown from the Shellflow collection of long-slit spectra. Smoothed versions of the observed velocity field were produced using the patch routine within the GIPSY analysis package (van der Hulst et al. 1992; Vogelaar & Terlouw 2001). The SparsePak velocity field shown in the lower window is extracted from a model which includes inclination, PA, disk center, rotation velocity, scale length, and systemic velocity. The parameterization of the velocity field is given by Eq. (1). SparsePak Hα position-velocity diagrams are constructed using 2 representations of the 2-D velocity field: The first includes all measurements with a simulated 6′′ "slit" for the best-fit kinematic position angle (filled triangles). The second SparsePak rotation curve uses all measured velocities within ±60 ◦ of the kinematic major axis in the inclined plane of the galaxy (open squares). Using the modeled kinematic inclination and position angle, we can project each measured rotation velocity onto the major axis. This second, "wedge," approach is relatively insensitive to inclination-induced beam smearing which affects the simulated slit measurements. However, the wedge does not spatially sample the inner 10′′ as well as the slit. Our best fit model (solid line) is adjusted for beam smearing induced by the ∼ 5′′ fibers of SparsePak. When comparing this model to the data, remember that the simulated slit data (filled triangles) have not been projected onto the major axis; the magnitude of these velocities serve only as a lower limit. Thus, a black triangle in the center of the RC that does not have a corresponding open box at the same radius implies that the center of that fiber lies more than 60◦ from the major axis and its azimuth correction is large (greater than 2). The velocity models based on Eq. (1) trace the open boxes only. Further details about velocity field modeling are given in §2.2. -- 19 -- REFERENCES Andersen, D. 2001, PhD thesis, Penn State University Andersen, D.R., Bershady, M.A., Sparke, L.S., Gallagher, J.S., Wilcots, E.M., van Driel, W., & Monnier-Ragaigne, D. 2003, in prep. Andersen, D.R. & Bershady, M.A. 2003, in prep. Athanassoula, L. 2003, MNRAS, in print Barnes, E.I. & Sellwood, J.A. 2003, AJ, 125, 1164 Bell, E. F. & de Jong, R. S. 2001, ApJ, 550, 212 Bershady, M. A., Andersen, D. R., Harker, J., & Ramsey, L. W. 2003a, in prep. Bershady, M. A., Andersen, D. R., & Verheijen, M. V. 2003b, in prep. Beauvais, C. & Bothun, G. 2001, ApJS, 136, 41 Bosma, A. 1996, ASP Conf. Ser. 91: IAU Colloq. 157: Barred Galaxies, 132 Bosma, A. 2002, ASP Conf. Ser. Vol. 273: The Dynamics, Structure, & History of Galaxies, eds. G.S. Da Costa & H. Jerjen, 223 Bottema, R. 1997, A&A, 328, 517 Brighenti, F. & Mathews, W. G. 1997, ApJ, 486, L83 Broeils, A. H. & Courteau, S. 1997, ASP Conf. Ser. 117: Dark and Visible Matter in Galaxies and Cosmological Implications, 74 Buchhorn, M. 1992, Ph.D. Thesis, Australian National University Buta, R., Crocker, D.A., & Elmegreen, B.G. 1996, ASP Conf. Ser. 91: Barred Galaxies Ciotti, L. 2000, ASP Conf. Ser. 197: Dynamics of Galaxies: from the Early Universe to the Present, 95 Corsini, E. M. et al. 1999, A&A, 342, 671 Courteau, S. 1996, ApJS, 103, 363 Courteau, S. 1997, AJ, 114, 2402 -- 20 -- Courteau, S. & Faber, S. M. 1988, PASP, 100, 1219 Courteau, S. & Rix, H. 1999, ApJ, 513, 561 Courteau, S., Willick, J. A., Strauss, M. A., Schlegel, D., & Postman, M. 2000, ApJ, 544, 636 Courteau, S., Dekel, A., & MacArthur, L.A. 2003a, in preparation Courteau, S. , Strauss, M. A., Schlegel, D., Postman, M., & MacArthur, L.A. 2003b, in preparation Dalcanton, J. J., Spergel, D. N., & Summers, F. J. 1997, ApJ, 482, 659 Dale, D. A., Giovanelli, R., Haynes, M. P., Campusano, L. E., & Hardy, E. 1999, AJ, 118, 1489 Debattista, V. P. & Sellwood, J. A. 2000, ApJ, 543, 704 Debattista, V. P. & Sellwood, J. A. 1998, ApJ, 493, L5 Dehnen, W. & Binney, J. 1998, MNRAS, 294, 429 Dutton, A., Courteau, S., & de Jong, R. 2003, in prep. Eskridge, P. B. et al. 2002, ApJS, 143, 73 Fuchs, B. 2001, Dark Matter in Astro- and Particle Physics, 25 Fuchs, B. 2002, [astro-ph/0209157] Genzel, R., Baker, A. J., Tacconi, L. J., Lutz, D., Cox, P., Guilloteau, S., & Omont, A. 2003, ApJ, 584, 633 Gerhard, O. 2002, in Matter in the Universe, eds. P. Jetzer, K. Pretzl, & R. von Steiger, Space Science Reviews (Kluwer), 100, 129 Giovanelli, R. & Haynes, M. P. 2002, ApJ, 571, L107 Jimenez, R., Verde, L., & Oh, S. P. 2003, MNRAS, 339, 243 Kennicutt, R.C. 1999, ARA&A, 36 Klypin, A., Zhao, H., & Somerville, R. S. 2002, ApJ, 573, 597 Kranz, T., Slyz, A., & Rix, H. 2003, ApJ, 586, 143 -- 21 -- Kuijken, K. 1995, in Stellar Populations, IAU 164 (Dordrecht: Kluwer), 195 MacArthur, L. A., Courteau. S., & Holtzman, J. A. 2003, ApJ, 582, 689 Maller, A. H., Simard, L., Guhathakurta, P., Hjorth, J., Jaunsen, A. O., Flores, R. A., & Primack, J. R. 2000, ApJ, 533, 194 Mo, H. J., Mao, S., & White, S. D. M. 1998, MNRAS, 295, 319 Ostriker, J. P. & Peebles, P. J. E. 1973, ApJ, 186, 467 Palunas, P. & Williams, T. B. 2000, AJ, 120, 2884 Peletier, R. F. & Willner, S. P. 1991, ApJ, 382, 382 Phillips, A. C. 1996, ASP Conf. Ser. 91: Barred Galaxies, eds. R.Buta, D. A. Crocker, & B. G. Elmegreen, 44 Press W. H., Teukolsky S. A., Vetterling W. T., & Flannery, B. P. 1992, Cambridge: Cam- bridge University Press) Sackett, P.D. 1997, ApJ, 483, 103 Sakai, S. et al. 2000, ApJ, 529, 698 Santiago, B. X., Strauss, M. A., Lahav, O., Davis, M., Dressler, A., & Huchra, J. P. 1995, ApJ, 446, 457 Seljak, U. 2002, MNRAS, 334, 797 Sellwood, J. A. 2003, ApJ, in print [astro-ph/0210079] Sellwood, J. A. & Evans, N. W. 2001, ApJ, 546, 176 Sheth, K., Vogel, S. N., Teuben, P. J., Harris, A. I., Regan, M. W., Thornley, M. D., & Helfer, T. T. 2002, ASP Conf. Ser. 275: Disks of Galaxies: Kinematics, Dynamics and Perturbations, 267 Sheth, K., Regan, M.W., Scoville, N.Z., & Strudde, L.E. 2003, ApJL, in print Silk, J. 2002, ASP Conf. Ser. 273: The Dynamics, Structure & History of Galaxies: A Workshop in Honour of Professor Ken Freeman, 299 Swaters, R. A. 1999, Ph.D. Thesis, Groningen University -- 22 -- Swaters, R. A., Madore, B. F., & Trewhella, M. 2000, ApJ, 531, L107 Swaters, R. A., M. A. W. Verheijen, M. A. Bershady, D. R. Andersen 2003, ApJ, 587, 19 Trott, C. M. & Webster, R. L. 2002, MNRAS, 334, 621 Tully, R. B. & Fisher, J. R. 1977, A&A, 54, 661 van den Bergh, S., Abraham, R. G., Whyte, L. F., Merrifield, M. R., Eskridge, P. B., Frogel, J. A., & Pogge, R. 2002, AJ, 123, 2913 van Dokkum, P. G. & Stanford, S. A. 2001, ApJ, 562, L35 van der Hulst, J. P. Terlouw, K. Begeman, W. Zwitser and P.R. Roelfsema, 1992, ASPCS, 25, 131 Valenzuela, O. & Klypin, A. 2003 (astro-ph/0204028) Vogelaar, M. G. R. & Terlouw, J. P. 2001, ASP Conf. Ser. 238: Astronomical Data Analysis Software and Systems X, eds. F. A. Primini, F.R. Harnden, Jr., H. E. Payne, 10, 358 Weinberg, M.D., & Katz, N. 2002, ApJ, 580, 627 Weiner, B. J., Sellwood, J. A., & Williams, T. B. 2001, ApJ, 546, 931 Zhao, H., Rich, R. M., & Spergel, D. N. 1996, MNRAS, 282, 175 This preprint was prepared with the AAS LATEX macros v5.0. -- 23 -- -- 24 -- Fig. 1. -- SparsePak fiber footprint for one pointing overlayed on our CCD I-band image for the SBbc galaxy UGC 5141. -- 25 -- Fig. 2. -- Differences in measurements of kinematic and photometric position angles and inclinations for galaxies with available 2-D velocity fields and I-band imaging. Inclinations shown next to the galaxy names correspond to the kinematic and photometric estimates, respectively. Inclination differences are larger for progressively face-on orientations. -- 26 -- 10 0 −10 −20 −30 10 0 −10 −20 −30 ) c i r t e m o t o h p − c i t a m e n k ( i i ∆ 0 20 40 ikinematic 60 80 Fig. 3. -- Difference between kinematic and photometric inclinations vs kinematic inclination for four galaxy samples. The point types are squares (this study); circles (Andersen 2001); triangles (Peletier & Willner 1991); and pentagons (Sakai et al. 2000). Open, gray-filled, and black-filled symbols represent unbarred, weakly-barred, and strongly barred galaxies, respectively. The top panel shows simple regressions to the Courteau et al. (this study), Andersen, and Peletier & Willner samples (independently). We exclude the Peletier & Willner sample in the bottom sample. -- 27 -- Fig. 4. -- Line width-luminosity (top) and size-luminosity (bottom) diagrams for Shellflow galaxies. Line widths are measured at 2.2 disk scale lengths and disk scale lengths are obtained from B/D decompositions of the surface brightness profile. Barred galaxies have filled symbols consistent with their Hubble type and are further emphasized with an open circle. Barred galaxies lie below the mean TFR, appearing to be systematically brighter for their rotational velocity. As in Sakai et al. (2000), this is a small number artifact. The solid line is a fit from our data−model minimization technique (Courteau et al. 2003a). -- 28 -- Fig. 5. -- Line width-luminosity (top) and size-luminosity (bottom) diagrams for SCII galax- ies. Line widths are measured from Hα rotation curves and HI line widths and disk scale lengths are measured using the "marking the disk" technique (see text). Symbols are as in Fig. 4. The TFR is the same for barred and unbarred galaxies. The solid and dashed lines show data−model minimization fits from Courteau et al. (2003a) and Dale et al. (1999), respectively. -- 29 -- Fig. 6. -- Color-magnitude diagram for Shellflow galaxies. Barred galaxies have mean colors consistent with the general spiral population. Symbols are as in Fig. 4. -- 30 -- IC 0784 60o SparsePak wedge 6" SparsePak slit SparsePak model Mean Velocity Error 250 200 150 100 50 0 −50 −100 −150 −200 ) s / m k ( t o r V −250 −80 −60 −40 0 −20 20 R (arcseconds) 40 60 80 Fig. 7. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for IC 784. -- 31 -- Fig. 8. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for IC 2104. This galaxy has a pathological velocity field with significant non-circular motions, a continuously rising rotation curve, and a small inner velocity bump representative of a strong bar and/or bulge (that is poorly matched by the velocity model). Fig. 9. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for NGC 2540. Fig. 10. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for the unbarred galaxy NGC 3029. Fig. 11. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for NGC 3128. Fig. 12. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for NGC 3469. Fig. 13. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for NGC 3832. Fig. 14. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for NGC 4999. Fig. 15. -- Velocity contours and V-band image (top) and rotation curve data with velocity model (bottom) for NGC 5504. The I-band image was not available. Fig. 16. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for UGC 4416. The vertical trace in the upper image is due to an internal image reflection. Fig. 17. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for UGC 5141. Fig. 18. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for the weakly barred galaxy UGC 6895. -- 32 -- Fig. 19. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for UGC 7173. Fig. 20. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for UGC 8229. Fig. 21. -- Velocity contours and I-band image (top) and rotation curve data with velocity model (bottom) for UGC 8241.
astro-ph/0308521
1
0308
2003-08-29T00:16:42
A Spectroscopic Technique for Measuring Stellar Properties of Pre-Main Sequence Stars
[ "astro-ph" ]
We describe a technique for deriving effective temperatures, surface gravities, rotation velocities, and radial velocities from high resolution near-IR spectra. The technique matches the observed near-IR spectra to spectra synthesized from model atmospheres. For pre-main sequence stars, we use the same matching process to also measure the amount of excess near-IR emission. The information derived from high resolution spectra comes from line shapes and the relative line strengths of closely spaced lines. The values for the stellar parameters we derive are therefore independent of those derived from low resolution spectroscopy and photometry. The new method offers the promise of improved accuracy in placing young stellar objects on evolutionary model tracks. We discuss the possible systematic effects on our determination of the stellar parameters and evaluate the accuracy of the results derivable from high resolution spectra. The analysis of high resolution near-IR spectra of MK standards shows that the technique gives very accurate values for the effective temperature. The biggest uncertainty in comparing our results with optical spectral typing of MK standards is in the spectral type to effective temperature conversion for the standards themselves. Even including this uncertainty, the 1 sigma difference between the optical and IR temperatures for 3000-5800 K dwarfs is only 140 K. In a companion paper (Doppmann, Jaffe, & White 2003), we present an analysis of heavily extincted young stellar objects rho Oph.
astro-ph
astro-ph
A Spectroscopic Technique for Measuring Stellar Properties of Accepted August 27, 2003 to AJ Pre -- Main Sequence Stars G. W. Doppmann1,2,3 and D. T. Jaffe2,3 [email protected] [email protected] ABSTRACT We describe a technique for deriving effective temperatures, surface gravi- ties, rotation velocities, and radial velocities from high resolution near -- IR spec- tra. The technique matches the observed near -- IR spectra to spectra synthesized from model atmospheres. Our analysis is geared toward characterizing heavily reddened pre -- main sequence stars but the technique also has potential applica- tions in characterizing main sequence and post -- main sequence stars when these lie behind thick clouds of interstellar dust. For the pre -- main sequence stars, we use the same matching process to measure the amount of excess near -- IR emis- sion (which may arise in the protostellar disks) in addition to the other stellar parameters. The information derived from high resolution spectra comes from line shapes and the relative line strengths of closely spaced lines. The values for the stellar parameters we derive are therefore independent of those derived from low resolution spectroscopy and photometry. The new method offers the promise of improved accuracy in placing young stellar objects on evolutionary model tracks. Tests with an artificial noisy spectrum with typical stellar param- eters, and signal -- to -- noise of 50 indicates a 1σ error of 100 K in Teff, 2 km s−1 in v sin i, and 0.13 in continuum veiling for an input veiling of 1. If the line flux ratio between the sum of the Na, Sc, and Si lines at 2.2 µm and the (2 -- 0) 12CO 1NASA Ames Research Center, MS 245-6, Moffett Field, CA 94035-1000 2Department of Astronomy, 1 University Station C1400, Austin, TX 78712-1083 3Visiting Astronomer, Kitt Peak National Observatory, National Optical Astronomy Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc. (AURA) under cooperative agreement with the National Science Foundation. -- 2 -- bandhead at 2.3 µm is known to an accuracy of 10%, the errors in our best fit value for log g will be ∆ log g = 0.1 -- 0.2. We discuss the possible systematic effects on our determination of the stellar parameters and evaluate the accu- racy of the results derivable from high resolution spectra. In the context of this evaluation, we explore quantitatively the degeneracy between temperature and gravity that has bedeviled efforts to type young stellar objects using low resolu- tion spectra. The analysis of high resolution near -- IR spectra of MK standards shows that the technique gives very accurate values for the effective temperature. The biggest uncertainty in comparing our results with optical spectral typing of MK standards is in the spectral type to effective temperature conversion for the standards themselves. Even including this uncertainty, the 1σ difference between the optical and IR temperatures for 3000 -- 5800 K dwarfs is only 140 K. In a companion paper (Doppmann, Jaffe, & White 2003), we present an analysis of heavily extincted young stellar objects in the ρ Ophiuchi molecular cloud. Subject headings: infrared: stars -- methods: analytical -- stars: fundamental parameters -- stars: techniques -- stars: spectroscopic 1. Introduction Our understanding of main sequence and post main sequence stellar evolution is a triumph of collaboration between theorists building physical models and observers collecting precise observations. Over many decades, the two groups have compared their results on the common ground of the Hertzsprung -- Russell (H -- R) diagram. For young stellar objects (YSOs) however, the link between theory and observation is much more tenuous. There is a well -- developed empirical scheme that classifies YSOs by their broad -- band spectral energy distributions (Lada 1987). Class I sources have rising near -- to mid -- IR spectra. Class II sources have broad, largely flat infrared spectral energy distributions, and Class III sources have spectral energy distributions of hot reddened blackbodies. Since the emission that distinguishes Class I and II objects from Class III sources does not arise from the stellar photosphere, models of pre -- main sequence (PMS) evolution are not strongly constrained by matching theoretical and observed spectral energy distributions. Comparison of model tracks with observational H -- R diagrams has been possible for visible T Tauri (Class II) stars (TTSs) (e.g. Huang 1961; Stahler 1988; Hatmann et al. 1991; Kenyon & Hartmann 1995; White et al. 1999; Webb et al. 1999; Simon et al. 2000; White & Ghez 2001). Obscuration by dust makes it difficult to study Class I objects and to observe TTSs in the visible if these objects are in clusters or associations within molecular clouds. The -- 3 -- resulting difficulties in comparing tracks with observations leave our understanding of the physics of YSOs on a much less solid footing than our understanding of more evolved objects. In studying YSOs, our main goal is to understand the history of these objects from their formation to their arrival on the main sequence. We also wish to use the ensemble of such objects to characterize forming clusters and associations; their initial mass functions and star formation histories. To meet these goals, we need to determine parameters of the stars and relate them to other properties of protostellar disks and of the surrounding cloud. The most important stellar parameters to obtain are the effective temperature and the luminosity or surface gravity since this pair permits us to place the YSOs into a theoretical H -- R diagram for comparison with theoretical PMS evolution models. (Note that luminosity and log g are not fully interchangeable since the relationship between these two quantities depends on the not yet well established mass -- luminosity -- radius relationship for YSOs.) Other useful parameters include the amount of reddening or extinction, the amount of non -- photospheric emission as a function of wavelength, and the stellar rotation rate. Observers have used many techniques to investigate the properties of stars in very young embedded clusters. In well -- studied regions like the ρ Ophiuchi cloud core, previous investigators have used photometric surveys in the near -- IR (Wilking & Lada 1983; Greene & Young 1992; Barsony et al. 1997), in the mid -- IR (Bontemps et al. 2001), and low -- to -- moderate resolution (R ≡ λ/∆λ = 500 -- 2000) near -- IR spectroscopy (Casali & Matthews 1992; Greene & Meyer 1995; Greene & Lada 1996; Kenyon et al. 1998; Luhman & Rieke 1999), to estimate temperatures, luminosities, and the amount of excess (non -- photospheric emission) in the infrared for the sources. In embedded clusters, however, even near -- IR photometry and low resolution spectroscopy suffer under disadvantages not inflicted upon these techniques when applied to main sequence stars or to less heavily extincted young stars. Problems include extremely high extinction (e.g. the central part of the ρ Ophiuchi molecular cloud where Av = 40 ± 10.9 magnitudes, Luhman & Rieke 1999), and excess emission in the near -- and mid -- IR (from warm dust in the circumstellar disks, Greene et al. 1994; Strom et al. 1995). Observers recognized long ago that spectral classification in the near -- IR was a poten- tially valuable tool for deriving the properties of obscured stars in the Galactic Plane and young stars obscured by their natal clouds (Merrill & Ridgway 1979). Young stars are brighter in the infrared both because their photospheres tend to be cool and because it is easier for the stellar infrared emission to penetrate through the dust within the star forming cloud or along the line -- of -- sight. Also, in TTSs, the ratio of photospheric flux to the hot continuum that produces the excess emission frequently seen in the UV (presumably from accretion shocks, Gullbring et al. 2000) is higher in the infrared than at shorter wavelengths, -- 4 -- permitting better detections of photospheric lines. The near -- IR spectra of late -- type stars contain useful information about the luminosities of the targets. The strength of the CO overtone bands at 2.3 µm was recognized early -- on as a useful indicator of luminosity, albeit with additional sensitivity to temperature (Baldwin et al. 1973; Kleinmann & Hall 1986; Lancon & Rocca -- Volmerange 1992). Ram´irez et al. (1997) find that an index formed from the equivalent widths of the strong near -- IR lines of neutral calcium and sodium and the (2 -- 0) 12CO bandhead is a luminosity indicator, independent of temperature, for giants in the range K0 to M6. Accurate estimates of spectral type are also possible from near -- IR spectra. Kleinmann & Hall (1986) calculated equivalent widths of key features from their K -- band spectra of MK standards and derived the dependence of these equivalent widths on spectral and luminosity class. Meyer et al. (1998) derived line ratio relations from H -- band spectra of MK standards and found that the relations agree with optical spectral types to within ±2 subclasses. They argue that using line ratios, rather than equivalent widths, makes the Teff determination less sensitive to the presence of continuum excesses in PMS objects. While most of the efforts to derive spectral types from the IR spectra have focused on empirical equivalent width or line ratio to Teff relations, there have also been a number of efforts to match low resolution H and K -- band spectra to synthetic spectra (Kirkpatrick et al. 1993; Ali et al. 1995; Leggett et al. 1996). Ali et al. (1995) find that they can match the temperatures of dwarfs with an error of ±350 K using this technique. All of the studies we have described so far were carried out with resolving powers below 3000, where the unsaturated photospheric features are unresolved. At these resolving powers, not only can the depths of the lines get quite small, especially in the presence of excess continuum emission but there can also be problems due to line blending. For example, many of these studies use the equivalent width of the Na features at 2.2 µm in the determination of spectral type and luminosity class. At R = 1, 000, the typical resolving power for the studies, however, the Na lines are blended with weaker but significant lines of Sc and Si that have very different dependences on Teff and log g. Existing spectra of the Na interval at a greater resolving power have shown that the relative strengths of the Sc, Si, and Na lines plus the shape of the Na features could potentially be sensitive indicators of effective temperature for dwarfs (Wallace & Hinkle 1996; Greene & Lada 1997). We present here a technique for deriving some of the properties of PMS stars from high resolution near -- IR spectra. High resolution spectra are particularly useful because the closely spaced lines (like the Na, Sc, and Si features at 2.206 µm, see Figures 1 & 2) in late -- type stars are observable separately and because it is possible to use the spectra to glean information from the shapes of the absorption lines (Johns -- Krull & Valenti 2001) -- 5 -- The stellar parameters we derive include the effective temperature (Teff), the surface gravity (log g), the rotation rate (v sin i), and the amount of stellar/circumstellar emission in excess of the photospheric emission (rλ). The method is largely independent of photometric data (except as used to cross -- calibrate high resolution spectral segments at different wavelengths and to estimate the small amount of reddening difference between 2.2 µm and 2.3µm, see Appendix A) and offers the promise of improved accuracy, especially for derivations of the properties of heavily obscured stars. In this paper, we present a quantitative derivation of the physical properties of PMS stars from high resolution spectra taken in the K -- band (2.0 -- 2.4 µm). Stars in the mass range from 0.1 to 0.9 M⊙ will lie in the spectral type range from M6 to K0 from the time they become visible in the near -- IR until they reach the main sequence (D'Antona & Mazzitelli 1997; Baraffe et al. 1998; Palla & Stahler 1999; Siess et al. 2000). For such late -- type stars and even for somewhat earlier types (later than G5), the near -- IR spectra are rich in lines of neutral metals and hardy molecules. It is therefore possible to use high resolution spectra from a limited spectral range to provide us with many of the important physical parameters of young or obscured stars. Our analysis involves a comparison of synthetic spectra to the observed high resolution data. The observations and data reduction are described in § 2. In § 3, we detail the basic spectral analysis technique with particular emphasis on the features of our method necessary to deal with the peculiarities of PMS objects. In § 4, we analyze the internal errors and investigate inherent systematic uncertainties. We compare results for standard stars from optical spectroscopy to the results we obtain through the analysis of high resolution near -- IR spectra, in § 5, and discuss the ways in which our derived properties supplement or improve upon properties measured with other techniques. In a companion paper (Doppmann, Jaffe, & White 2003, hereafter DJW03), we present and analyze near -- IR spectra of a sample of Class II YSOs in the core of the ρ Ophiuchi molecular cloud. 2. Observations and Data Reduction In order to test our technique for derivation of stellar parameters from high resolution spectra in the 2.2 µm atmospheric window, we have assembled a sample of spectra of MK standards. Table 1 lists the stars in this sample, the instruments used to obtain the data, and various stellar properties obtained from the literature. We observed part of the sample using the PHOENIX spectrograph (Hinkle et al. 1998), on the Kitt Peak 4 meter in May 2000. The observed spectral interval covered 95 A cen- tered at 2.2070 µm. The resolving power for these observations was R ≡ λ/∆λ ∼= 50,000. Individual pixels covered a range in wavelength of λ/∆λ = 240,000. -- 6 -- A partially overlapping set of MK standards was observed using the NIRSPEC instru- ment (McLean et al. 1995), on the Keck Telescope in May 2000. These spectra were kindly provided to us by Tom Greene. For these observations, NIRSPEC had a slit -- limited resolv- ing power of 17,500 and covered a total of 230 nm over 6 non -- contiguous orders in the 2 µm atmospheric window. Observations with both instruments were made by nodding the telescope, placing the target star alternately at two different positions along the slit. IRAF was used to reduce the spectra from both spectrometers in roughly same way (for details about the NIRSPEC data reduction see Greene & Lada 2000). With the PHOENIX data, we differenced the source frames taken at alternate slit positions and divided by flat fields using an internal continuum lamp that uniformly illuminated the entire slit. At loca- tions of bad pixels, we substituted values interpolated from neighboring positions. We then optimally extracted the spectra using IRAF's apall package. We used telluric absorption lines of H2O and CH4 to wavelength calibrate the PHOENIX data. The best fit to the low order wavelength solution was accurate to 0.1 pixels along the detector array. We removed the telluric lines from the spectra by dividing the data by spectra of early -- type stars taken at the same airmass (typically within 5o of the target). We produced the final spectra by taking a signal -- weighted average of the calibrated spectra from the two beam positions. To prepare the final spectra for comparison with models, we set a continuum level by eye using regions in the observed spectra where synthesis models indicate that strong lines and extended line wings are not present. 3. Method Ideally, we would like to be able to derive stellar parameters across the whole range of masses and ages present in clusters or associations of newly forming stars. The actual range of surface gravities and effective temperatures for which we can derive parameters from high resolution spectroscopic observations using the technique we outline here is constrained in several ways: we are limited to objects that permit observable amounts of near -- IR radiation from the stellar photosphere to escape through the surrounding disk and envelope. In general, photometric studies of YSOs imply that the lowest surface gravities for which objects become visible in the near -- IR are log g ≈ 3.0 (Comer´on et al 1993; Siess et al. 2000). Our technique requires that the stellar photospheres have a sufficient number of reasonably strong lines and that these lines be sensitive to variations in temperature and surface gravity. Further, it requires that the stellar models and available line lists be adequate to permit us to make accurate high resolution spectral syntheses for comparison with the observed spectra. With the stellar atmospheres and synthesis program we are using, our analysis technique works -- 7 -- well within the range 3000 ≤ Teff ≤ 5800 K, and 3.5 ≤ log g ≤ 5.5. Since H -- R diagram evolutionary tracks for low mass PMS objects are largely vertical, the temperature constraint implies a range of masses for which we can position objects in the H -- R diagram of roughly 0.1 to 1.6 M⊙. 3.1. Technique Overview The basis for our technique is a grid of synthetic high resolution spectra in the K window (2.0 -- 2.4 µm). The grid spans the relevant ranges of the important stellar parameters for YSOs: effective temperature, surface gravity, veiling, and v sin i rotation. The best model fit is chosen by an RMS minimization of the residuals across the photospheric absorption lines in our spectral window. This minimization also includes a fit for the stellar radial velocity. The 2.0 -- 2.4 µm atmospheric window contains features that are sensitive to both the temperature and pressure in the stellar photosphere. It is also the longest wavelength band where the photospheric emission from the youngest stars is comparable in flux to the thermal emission from dust in the circumstellar disk. At this wavelength, the sensitivity of ground -- based spectrometers with large resolving powers is not yet compromised by thermal emission from the telescope and sky. The 2 µm band is also not far from the maximum in the photospheric emission from late -- type stars. For the purposes of our spectral matching program, high spectral resolution means sufficient resolution to permit us to resolve most stellar lines (R ≥ 20,000). No existing high resolution spectrometer covers the entire 2.0 -- 2.4 µm atmospheric window with a single exposure. The instrument with the most coverage (NIRSPEC, McLean et al. 1995) gives cross -- dispersed spectra covering about 1/3 of the window in disconnected segments. Other existing instruments cover only individual 50 -- 200 Angstrom bands (PHOENIX, Hinkle et al. 1998), (CSHELL, Greene et al. 1993), (CGS4, Mountain et al. 1990; Wright et al. 1993). In order to be generally applicable without enormous expenditure of telescope time, our technique should therefore use only a limited part of the spectrum available within the atmospheric window. We have used spectral synthesis models to explore the K window to find the spectral in- tervals that have strong lines that vary significantly with variations in the stellar parameters. Based on this investigation, we chose the region at 2.2070 µm ±0.0050 (hereafter "the Na interval", see Figure 1a) which includes two strong neutral sodium lines (4s2S1/2 -- 4p2P0 3/2), and several prominent lines of neutral Si and Sc. We selected a second spectral region at 2.2960 µm ±0.0035 where the dominant feature is the 2 -- 0 bandhead of 12CO ("the 12CO -- 8 -- interval", see Figure 1b) for our analysis. Figure 2 presents a sequence of spectral syntheses for the Na interval that illustrates the sensitivity of the line ratios and line shapes in this wavelength band to the photospheric temperature. We produced a grid of spectra covering the appropriate range in Teff and log g using the NextGen non -- grey atmosphere models (Hauschildt et al. 1999). These models include TiO and H2O opacities, critical for cooler atmospheres. We synthesized a high resolution (R = 120, 000) K -- band spectrum at 2.2 µm and 2.3 µm using the MOOG spectral synthesis code (Sneden 1973). Atomic and CO line lists came from Kurucz (1994) and Goorvitch & Chackerian (1994), respectively. We have computed all models with solar metallicity. The critical relative abundance in our analysis is [Sc/Si]. Stellar abundances of [Si/Fe] in the local neighborhood are solar (Edvardsson et al. 1993), and we assume the same for [Sc/Fe]. For our analysis, we use synthesis models computed with solar microturbulence values (1 km s−1). Gray, Graham, & Hoyt (2001) have compared spectral synthesis models to optical spectra of MK standards deriving values of Teff, log g, and microturbulence for stars with spectral types from A5 to G2 and log gs from 1.2 to 4.5. At all temperatures, they see a trend in the best -- fit value of the mircoturbulence. This value decreases as log g increases, approaching a roughly constant value at log g > 3.5. This asymptotic value of log g decreases steadily from ∼2.5 km s−1 at spectral type A6 to ∼1.3 km s−1 at spectral type G1, lending further support to the appropriateness of our use of the solar microturbulence value in our models of late -- type stars with dwarf and sub -- dwarf gravities. At cool temperatures (i.e. Teff < 4000 K), the wings of the two Na I lines are noticeably pressure broadened. The synthesis code, MOOG, allows for the van der Waals damping parameter to be adjusted in creating the artificial spectra. We have tuned the amount of damping present in the Na I lines to give the best fit to an observed spectrum of the sun (Livingston & Wallace 1991). The best fit was the Unsold (1950) approximation used to calculate the van der Waals damping constant. The strength of the Na lines increases with increasing surface gravity at a fixed temper- ature, while CO lines decrease in strength. A qualitative way to understand these trends is to examine how the increase in electron presssure drives where the lines and continuum form within the stellar atmosphere. The greater electron pressures at higher gravity results in a larger fractional abundance of neutral sodium, causing the 2.2 µm lines to form closer to the stellar surface. This effect is larger than the decrease in the depth of continuum formation with increasing electron pressure resulting in larger line depths for Na. In the case of CO, the increase in the continuum opacity of H− is the dominant effect and the continuum layer moves closer to the line forming region reducing the strength of the bandhead absorption. -- 9 -- We also wish to derive the rotation rate (v sin i), and the amount of "veiling" (non -- photospheric excess emission relative to the photospheric flux). At a given wavelength, for example that of the Na interval, this veiling is defined as rNa ≡ (Fsource−Fphot)/Fphot, where Fsource and Fphot are the observed flux and the photospheric flux, respectively. Therefore, we added extra dimensions to the Teff -- log g grid of synthetic spectra to include variations in v sin i and rNa as well. To account for v sin i variations, we convolved the spectra with a rotational broadening profile that had an assumed limb darkening coefficient of 0.6 (Gray 1992) adding rotation at rates ranging from v sin i = 1 to 40 km s−1. The intensity of the non -- photospheric emission over the individual narrow spectral intervals has at most a slope of a few percent, so we can treat it with a single parameter. In the presence of such an excess, resolved lines have lower equivalent widths but retain the shapes imparted to them by line transfer in the stellar atmosphere and by rotation. We alter the synthetic spectra to include additional continuum to simulate veiling (rNa) ranging from 0 to 8. 3.2. Actual technique While, in all cases, our technique involves precise matching of spectra synthesized from model atmospheres to high resolution observations of limited portions of stellar spectra in the near -- IR, the exact procedure and which parameters we can derive depend on the nature of the target stars and on the data available. The interval around the 2.2 µm sodium lines is particularly rich in diagnostic power. When observations of only this interval are available, we can determine both the size of any excess emission and the rotation velocity in PMS stars, as well as a quite well constrained measure of the stellar effective temperature, all without recourse to stellar photometry that is sensitive to reddening and to the circumstellar excess. We discuss our analysis for stars where only the Na interval has been observed in § 3.2.1. With the addition of observations of the 12CO interval, it is possible to refine the Teff determination and to determine the surface gravity of the emitting star, even in the presence of significant veiling and reddening. In § 3.2.2, we explain how these improvements come about. Throughout the discussion, our focus is on applying the technique to reddened young TTSs (DJW03). When using this method to characterize heavily reddened main sequence stars, only minor modifications (such as dropping the free parameter for near -- IR veiling) are needed. -- 10 -- 3.2.1. Obtaining Teff, v sin i, and rNa from the Na Interval Alone We begin the spectral matching for the Na interval by using pattern recognition to constrain the effective temperature to a 1000 K range. For YSOs, we fix the value of log g =3.5, which corresponds to ∼1 -- 2 Myr old objects in stellar evolutionary models (Baraffe et al. 1998; Siess et al. 2000; Palla & Stahler 2000), consistent with age estimates of the central embedded cluster in Ophiuchus that includes the sources studied in DJW03 (Wilking et al. 1989; Greene & Meyer 1995; Luhman & Rieke 1999; Bontemps et al. 2001). In § 4 below, we describe in detail how the choice of a particular value of log g affects the Teff determination and how to correct Teff should log g have a different value. Once we have the grid of spectral syntheses in place, we need to choose spectral subin- tervals over which to compare the spectra synthesized from the atmosphere models to the observed spectra. Based on our experience with both MK standards and YSOs, we restrict the subintervals to the regions within the observed spectra where there is measurable line absorption. The upper and lower boundaries of the wavelength range vary with the apparent width of the stronger spectral features. The next step is to take the continuum normalized observed spectrum and compare it to each synthesized model spectrum calculating the RMS difference over the subintervals chosen in the previous step. Figure 3 illustrates the minimization by showing an artificial noisy spectrum and how the differences between this spectrum and the noiseless synthetic spectra vary as the search routine steps through the correct value of Teff. We then find the combination of Teff, rNa, and v sin i values that would produce the minimum RMS difference between the model and the observed spectra by interpolation. We illustrate the minimization process in Figure 4 which shows the variation of the RMS difference between an artificial noisy spectrum and various synthesis models of the Na interval as a function of the search parameters. We show variations of log g even though we would normally fix this parameter at a best -- guess value when only Na interval data are available. The figure displays the RMS difference and thereby the shape of the minimum in the Teff -- v sin i plane, the Teff -- rNa plane, and the Teff -- log g plane as the two variables are varied while the other two variables are held fixed at their nominal values. In all three planes, there are well -- determined minima in the fits to the spectrum of the Na interval for all variables except log g. The minimum in the Teff -- log g plane is very shallow. It is the one plane where we do not usually recover the input model. The grid in the log g direction is fairly coarse (∆ log g = 0.5) and some noise seeds for the artificial spectrum at log g = 4.0 even find the lowest RMS value at the edge of the grid (log g = 3.5), leaving the exact location of the minimum uncertain. This cut (see bottom panel of Figure 4) illustrates that an incorrect guess of log g for real target spectra can lead to systematic errors in the derived -- 11 -- value of Teff (see § 4). 3.2.2. Solving Simultaneously for Teff and log g There is a strong inverse dependence of the (2 -- 0) 12CO bandhead equivalent width on log g. There is also a weaker but noticeable dependence of the line equivalent width for the Na interval with log g. Figure 5 plots the ratio of (2 -- 0) 12CO to Na interval photospheric equivalent width (the equivalent width summed over the lines within the interval after re- moval of any non -- photospheric continuum from the spectrum (see Appendix A) as a function of Teff for log g values ranging from 3.5 to 5.0. We derived this ratio from the NextGen pho- tospheric models by creating synthetic spectra for the relevant intervals and integrating the spectra over the bands marked in Figure 1. For Teff > 3700 K, the ratio varies strongly with log g and is almost independent of temperature. For lower temperatures, the ratio still varies strongly with log g but Teff must also be known to correct for the sensitivity of the equivalent width ratio to temperature. In § 3.2.1, we derived Teff , v sin i, and rNa from spectra of the Na interval while holding log g fixed. When we have data available for both the Na and the 12CO intervals, we are able to solve for log g rather than just assume its value. We determine log g and the other parameters iteratively. We begin by assuming log g = 3.5 and derive a best fit for Teff , rNa, and v sin i using the Na observations as in § 3.2.1. We take the values of Teff and rNa that this process produces and use them, together with the integrals over the spectral intervals shown in Figure 1, to compute the ratio of photospheric equivalent widths in the 12CO and Na intervals (see Appendix A). With the photospheric equivalent width ratio and the derived value of Teff, we can use the relations plotted in Figure 5 to estimate log g and use this new value of log g in an iteration of the procedure for deriving Teff, v sin i and rNa from the observed spectrum of the Na interval. The iterative procedure converges quickly on a value for log g. The first two panels of Figure 6 illustrate how the derivation of parameters from the Na interval plus the use of the 12CO interval photospheric equivalent width work together to produce the correct value for all four parameters. The figure shows how the equivalent width of the CO bandhead varies significantly more than the equivalent width of the Na interval going from (Teff, log g) = (3600 K, 3.5) to (4000 K, 4.5). Note that at low temperatures and high surface gravities, the Na line wings extend beyond the spectral interval over which we compute the equivalent width for the Na interval. When the data cover a similar or narrower spectral interval, observations of cool and/or high surface gravity stars must be corrected for the fact that the intensity at the edges of the band is not fully at the level of the photospheric continuum. We have applied this correction in Figure 5. Therefore even when data with -- 12 -- broader spectral coverage are available, the measured equivalent width should be computed only over the marked intervals in Figure 1 for comparison to Figure 5. Although our procedure for using the CO interval together with the Na interval to determine log g does not, in principle, require high spectral resolution observations of the CO bandhead, such observations are very useful. In the youngest YSOs, the excess non -- photospheric emission can be many times greater than the emission from the photosphere itself. In such cases, the equivalent widths of the lines not only fail to represent the conditions in the stellar atmosphere but also can be extremely hard to measure. In low resolution spectra, the combination of dilution by non -- photospheric emission and dilution because the features are unresolved can make reliable measures of equivalent widths very problematic. Good atmospheric cancellation is also difficult because of the inherent messiness both of the stellar and the telluric spectrum in the region of the (2 -- 0) 12CO bandhead. Higher resolution spectra improve the situation because they make it easier to cancel telluric lines and because line -- to -- continuum ratios are larger. At high resolution, where we can resolve the CO bandhead and also the adjacent features in the ascending and descending R -- branch and where effective telluric line cancellation is possible, we can match the depth and shape of a CO bandhead to obtain useful information about both Teff and log g that does not depend strongly on precise determination of the continuum level. The third panel in Figure 6 illustrates the dependence of the spectral shape in the (2 -- 0) 12CO bandhead on log g and Teff as it would be seen at high spectral resolution. This panel demonstrates the potential for using detailed spectral shapes in the 12CO interval to improve the robustness of the stellar parameter derivation scheme. It shows the difference between the synthetic spectrum for the CO interval for (Teff, log g)=(3800 K, 4.0) and (4000 K, 4.5) and (3600 K, 3.5). We see that, in addition to the equivalent width changes, there is a change in the individual line depths and in the shape of the envelope of the bandhead at this resolving power (R = 50, 000). In the current work, however, we restrict ourselves to analyzing the results of detailed spectral synthesis matching for the Na interval combined with use of the Na/12CO photospheric equivalent width ratio to arrive at accurate values for Teff, v sin i, rNa and log g. 4. Analysis of Errors Now that we have developed a procedure for determining stellar parameters from high resolution spectra, we would like to know how well it works, both the sensitivity of the -- 13 -- fitting routine and the uniqueness of the derived solutions. We discuss here four classes of uncertainty: (1) The sensitivity of the model fitting to random noise in the spectra. (2) In- ternal systematic uncertainties arising from the optimization scheme and from degeneracies between various stellar parameters. (3) External systematic errors introduced by transfor- mations from one theoretical framework to another. (4) Errors arising from non -- random effects present in the data. We evaluate the effects of these problems on derived parameters using simulations, real data, and modifications of real data. We first analyze the errors in parameters derived using only the Na interval and then discuss changes in these errors and the uncertainty in the determination of log g when observations of both the Na and 12CO intervals are available. 4.1. Random Errors A generalized assessment of the sensitivity of our fitting routines to random errors is not possible. The sensitivity of the Teff , v sin i, and rNa determinations will be not only a function of the signal -- to -- noise ratio (S/N) of the spectra but also will depend on the widths of the lines and on their strengths relative to the continuum (that is, effectively on all four parameters). By working with a typical spectrum, however, we can get a rough idea of what S/N we require to reach the point where random noise in the spectra no longer dominates the uncertainty in determining stellar parameters. We illustrate the sensitivity to random errors by taking a synthetic spectrum for the model shown in Figure 1a and adding more and more Gaussian random noise to it. We take each artificially noisy spectrum, find the best -- fit values for Teff, v sin i and rNa in the usual way, by calculating the RMS difference between this spectrum and a set of noise -- free synthesis models covering a range in all three parameters and then interpolating to obtain the best -- fit values. At each S/N level, we re -- seed the noisy spectrum 30 times and repeat the fitting procedure. The standard deviation about the mean value of each derived parameter for this ensemble of noisy spectra then reflects the uncertainty due to random noise at a given S/N level. Figure 7 shows how the standard deviation about the mean derived Teff, v sin i, and rNa varies for noisy versions of the Na interval spectrum in Figure 1a (with continuum added to make rNa = 1) as the S/N decreases. For this particular case, the uncertainty in the Teff due to random errors is less than 100 K for S/N > 50. Random errors result in uncertainties of less than one spectral subclass as soon as the S/N ratio is greater than 35 at R = 50, 000. At S/N = 50, the random uncertainty in v sin i is 2 km s−1 and that in rNa is 0.13. The curves shown in Figure 7 illustrate the behavior of the random errors with S/N for a spectrum with rNa=1. For sources with other values of rNa, we can derive the S/N required for a given uncertainty in Teff , v sin i, or (1+rNa) by multiplying the value shown in Figure 7 by (1+rNa)/2. For a -- 14 -- given uncertainty in (1+rNa), the uncertainty in rNa itself is (1+rNa) times larger. Once the S/N exceeds ∼35×(1+rNa), other forms of errors begin to dominate the uncertainty in the derivation of stellar parameters from the K -- band spectral fitting technique. For v sin i, we also tested the dependence of the uncertainty on Teff. We used models with stellar rotation (25 km s−1), instrumental smoothing (R = 50, 000) and Gaussian noise (S/N = 30 per R = 240, 000 channel with rNa=0) added to simulate real data. We fit the modified synthetic spectra to a set of noiseless models keeping the temperature fixed at the correct value and allowing the RMS minimization algorithm to select the best v sin i value. This was repeated with 30 noise seeds for each temperature giving a mean and 1 σ error for v sin i for temperatures between 3200 K and 4400 K. The 1σ error is less than 2 km s−1 over the entire temperature range. 4.2. Uncertainties Arising from Internal Systematics For some pairs of stellar parameters, changes of both parameters simultaneously in a certain sense can keep the line shapes and depths almost unchanged. These partial degen- eracies in the output line shapes between different groupings of stellar parameters can have the effect of exaggerating the uncertainties caused by random errors. For minimizations of model -- observed spectral differences in the Na interval, this effect is most evident in the broadness of the minimum RMS error along a diagonal in the Teff − log g plane (Figure 4, bottom). When the spectra are noisy or imperfect, there is a range of temperature -- gravity pairs with very similar RMS values. Similar degeneracies for other lines in the near -- IR have caused difficulties when attempting to type YSOs from low resolution near -- IR spectra, un- less one knows what luminosity class is appropriate for the template stars (Luhman & Rieke 1999). The similar line shapes for models along a diagonal in the temperature -- gravity plane mean that we can introduce systematic errors in the derived Teff when we assume a value of log g and then derive the temperature. We can illustrate and assess this effect with fits to models. Figure 8 summarizes the results of our tests by showing how the derived temperatures deviate from the target spectrum temperature at different target values of Teff. Typically, our best fits for Teff using models with a log g differing by ±0.5 from the log g of the target spectra mis -- estimate the temperature by about 7%. This difference is approximately 1 -- 2 spectral subclasses over the range of Teff relevant to our study. If we later obtain an independent estimate of log g, we can correct the derived Teff. For data where we derive an effective temperature from the Na interval assuming some value of log g, we can correct the derived log Teff by +0.06 for every log g = 1 difference between the assumed -- 15 -- and the correct values. Making this correction, we recover the actual value of Teff to within better than 4%. The way in which we create the error space from comparisons of observed target spectra and synthesized model spectra may have a systematic effect on the derived parameters. The error space shown in Figure 4 represents the RMS deviation of the target spectrum from the models over selected intervals where line absorption was present (see Figure 3). This scheme allows for variations of feature depths and shapes. Because it is a straight RMS, it weights the stronger features, in particular the Na features, more heavily. One might ask if this is the best scheme, i.e. does it make the uncertainties in the derived parameters larger than they need to be? In its favor is the ability of the Na line depths to distinguish the value of rNa and the sensitivity of the depth and shape of these lines to Teff and log g. On the other side of the ledger is the low weight a straight RMS gives to the weaker Sc and Si lines. The ratio of these lines is the most sensitive temperature indicator in the Na interval. More complex schemes that make better use of the information content of the weaker lines and the subtleties of the line shapes are certainly possible. When the S/N gets large enough that systematic errors dominate over random errors, a straight RMS does not do as well as a weighted error scheme, but remains a reasonable approximation. The simple RMS scheme, however, has the great advantage that it finds the right parameters robustly over our whole temperature range at various values of rNa and v sin i and that its level of complexity is appropriate to S/N ratios of 30 -- 50 where random errors are just beginning to give way to systematics. 4.3. Errors in the Radial Velocity Determination When we work with real data, we use the RMS minimization of the difference between the data and the synthesis models to determine the best radial velocity shift of the observed stars along with the best -- fit stellar parameters. A narrow search range in radial velocity space is first selected by inspection. The data and models are interpolated to a higher dispersion (Rpix = 360, 000). We then use the minimum RMS of the residuals to the fit to select the best sub -- pixel radial velocity shift. Tests with artificial spectra show that random errors in the radial velocity determination due to noise in the spectrum are small (< 0.5 km s−1) for spectra with S/N > 30 per pixel (Rpix = 240, 000). We also examined how errors in the radial velocity fit affect our determination of the stellar parameters. For a S/N ∼30 per pixel, noise in the spectra is a much more significant factor in causing errors in v sin i than the error in the radial velocity determination. Stellar parameters determined mostly from line depths (i.e. effective temperature and veiling), are not sensitive to radial velocity errors at the 5 km s−1 level. -- 16 -- 5. Comparison with Standards and External Systematics We can use the spectra we have taken in the 2.2 µm Na interval of MK standard stars to perform a real -- world test of our fitting technique. We discuss here the test results for Teff and v sin i. We also add an artificial infrared excess and rotation velocity to the MK standards in order to look for systematic effects in the determination of rNa and refine our understanding of such effects on our determination of v sin i. The high resolution spectra we use to derive stellar parameters are imperfect representa- tions of the source spectra. Systematic problems with the data include imperfect flat -- fielding, defects in the cancellation of telluric features and the presence of scattered light or leaked out -- of -- order stellar radiation. The use of high resolution spectra in our analysis reduces the effect of these problems substantially. By choosing restricted wavelength intervals over which to match the synthetic models to the data, we minimize flat -- fielding effects since mismatches in the shapes of the synthetic and observed lines then play a bigger role in influencing the RMS difference. At high spectral resolution, the residuals of telluric lines cover a limited and known part of wavelength space. We simply exclude these regions from our fitting spectral matching intervals. Scattered light is usually removed in the data analysis if the source and sky have been switched regularly between two positions along the slit. The continuum level for our spectra could be another free parameter in our spectral fitting routine, though we have chosen to leave the continuum fixed when finding the mini- mum RMS difference between the target and synthetic spectra. For our target spectra, we have normalized the continuum based on a linear fit of two points close to the edges of the spectrum and farthest from Na I lines where the potential influence by any damping wings is minimized. In some cases where the instantaneous spectral coverage of the spectrometer is limited, it can be difficult to set the continuum level in the spectra correctly, particularly for the cooler stars and stars with higher surface gravities where the wings of the Na lines are extended. Differences between the effective temperatures derived from optical spectral types and the Teff values we derive from high resolution observations of the Na interval reflect the sum of the internal errors in our determination of Teff, errors in spectral typing from optical observations, and errors in converting from spectral types to effective temperatures. This last error can be substantial. De Jager & Nieuwenhuijzen (1987) estimate errors of 0.021 in the log of Teff in their spectral type -- Teff conversion for dwarfs, corresponding to ±200 K at -- 17 -- Teff=4000 K. Table 1 lists the effective temperature we derive for MK standards in our sample using the Na interval, as well as temperatures derived from the optical spectral types using the Teff -- spectral type relation of de Jager & Nieuwenhuijzen (1987). In Figure 9, we plot the near -- IR Teff against Teff determined from optical measurements by a variety of different techniques. For a given target, the spread of different symbols along the horizontal axis illustrates differences in the conversion between spectral type and temperature (de Jager & Nieuwenhuijzen 1987; Ali et al. 1995; Alonso et al. 1996; Allen 2000). The figure also includes temperatures derived using observed B -- V colors and the conversion relation of Kenyon & Hartmann (1995) which we will also apply to YSOs. The vertical error bars (placed on the de Jager Teff points) show our estimate of the 1σ uncertainty due to random noise in our data. The errors for our Teff determinations were derived by carrying out an analysis similar to that used to create Figure 7 but at the temperature and S/N ratio appropriate to each MK standard. The size of these errors indicates that systematic effects must dominate any differences between the optical and infrared results. For the luminosity class V MK standards, a best fit to the Teff (optical, de Jager & Nieuwenhuijzen 1987) versus Teff(Na interval) yields: Teff(Na) = 1.08 × Teff(de Jager) − 295 (1) The 1σ deviation from this relation is 113 K. For the same sources, the 1σ deviation for an "assumed" relation of Teff(Na)=Teff(de Jager) is 141 K, which is still less than the quoted uncertainty in the spectral type -- temperature conversion. All of the MK standards in our sample are fairly slow rotators and therefore not partic- ularly useful for tests of the accuracy of our fits to the Na interval for v sin i. Table 1 lists the values of v sin i from optical spectra as well as the results of our fit to the Na interval. For these near -- IR results, we have removed the effect of slit broadening by subtracting it in quadrature from the fitted value to produce the intrinsic widths or upper limits listed in the Table. For 10 of the 11 stars, the optical and near -- IR measurements and upper limits are consistent with each other. For the one discrepant source, HD 131976, where we measure v sin i= 10 km s−1, versus the value of 1.4 km s−1 measured by Duquennoy & Mayor (1988), the S/N for the infrared measurement is lower than that of any other standard in our sample. When we study a particular embedded PMS star, we usually do not know its metallicity. Assuming solar metallicity, however, should not cause problems for comparisons between models and stellar spectra since the deviations from solar metallicity are usually quite small, typically ∆[Fe/H] < 0.1 (Padgett 1996). For the MK standards in the field, differences -- 18 -- between the assumed and actual stellar abundance can cause systematic differences in the Teff scale. Eight of the MK standards we have observed as a test sample have measured abundances ranging from +0.02 dex above solar to −0.30 dex below (Table 1). For the comparison with MK standards in Figure 9, we fixed rNa=0 and used the metallicities listed in Table 1 or solar metallicities when no measurements were available. For MK standards with measured non -- solar metallicities, we computed separate grids of synthetic spectra for comparison with the observed spectra. Since the stellar atmosphere models available to us were gridded rather coarsely in metallicity for our purposes (every 0.5 dex), we constructed the spectral synthesis grids for mildly metal -- poor stars (0.3 ≤ [Fe/H] ≤ 0) using solar metallicity model atmospheres but synthesizing the spectra using abundances scaled down by [Fe/H]. This procedure does not account perfectly for metallicity effects. If we constrain the search grid to a fixed veiling (rNa = 0) and metallicity determined from the literature, however, the best fit model (as determined by the minimum RMS difference) has a noticeably non -- zero residual equivalent width when compared to the observed spectrum of the MK standard. In order to test the ability of our fitting routine to derive v sin i and rNa as well as Teff from data containing realistic amounts of systematic deviation from ideal spectra, we have altered our PHOENIX observations of MK standards (§ 2). We began with MK standards with very small rotation velocities (v sin i ≤ 10 km s−1). These stars also had no intrinsic veiling (rNa=0). As we showed above, if we hold rNa fixed at zero, we recover the Teff derived from optical spectra from our near -- IR observations. To each spectrum, we then added a known amount of rotation (v sin i = 25 km s−1) and veiling (tripling the amount of continuum to make rNa = 2.0). The Na interval spectra of these doctored stars were then analyzed using our standard procedure. For 6 of the 7 objects, we recovered effective temperatures close to those derived from the unaltered stellar spectra, typically within one spectral subclass of the best -- fit temperature for the unaltered spectrum (Table 2). The recovered temperature tended, however, to be systematically lower than the values derived holding rNa fixed and the recovered rNa values were higher than those we put into the spectra. The seventh object, HD 117176, has a Teff higher than the range for which the technique is fully reliable. For this source, the fitting routine derived a lower temperature and higher veiling. In all cases, the recovered v sin i was greater than or equal to but always within 5 km s−1 of the 25 km s−1 we put into the spectra. Table 2 also lists the values of rNa derived for the MK standards to which we had artificially added an rNa=2.0. As in the case of Teff, the fit for HD 117176 differs strongly from the input value. For the remaining luminosity class V sources, the derived rNa is 2.7 ±0.22. For the two luminosity class IV sources, the average value is 2.4. One possible contributor to this systematic difference may be the damping parameter used in the line -- 19 -- synthesis. The parameter that works for the solar spectrum may not be ideal for the cooler target stars. For the YSOs, whose gravities are more comparable to the luminosity class IV MK standards, this systematic problem may lead to an overestimate of rNa by ∆rNa= 0.13×(1+ rNa). Differences between the value of rNa derived from analysis of the Na interval spectra and values derived by other methods that are smaller than this ∆rNa are probably not significant. Further analysis with a better sample of MK standards will be needed to understand this effect more fully. On a real sample of PMS stars, one would be forced to address the effects of possible binary companions. Close companions will be recognizable in high resolution spectra because of their effect on stellar radial velocities. Beyond a few AU, the radial velocity effects will be much less apparent and imaging will be needed. A recent compilation of multiplicity data yields a companion star frequency for TTSs in nearby star forming regions of 24% ±11% (Ophiuchus) and 37% ±9% (Taurus), (Barsony, Koresko, & Matthews 2003), for bright companions (∆K ≤ 2 -- 3 mag). Of these, 40% have large enough separations to make them readily separable for direct imaging or spectroscopy. Therefore, ∼20% of a sample of stars in nearby star forming regions will have spectra that suffer from contamination of a secondary component. Spectroscopy using adaptive optics will eliminate this problem for the vast majority of sources. Continuum opacity due to molecules not present in our synthesis models is also a likely contributor to the systematic difference in the derived veiling values. Preliminary tests with numerous CO, SiO, OH, and H2O lines that blanket the Na interval reveal a ∼5% decrease in the continuum compared to the continuum determined from our synthesis models (Carbon 2003, private communication). Neglecting this decrease in the continuum relative to the cores of the photospheric lines could cause overestimates in the derived veiling by ∼0.05, as well as slightly altering the shapes of the atomic line wings. 6. Conclusions We have described, demonstrated and evaluated a technique for deriving effective tem- peratures, surface gravities, rotation rates, and infrared excesses from high resolution spectra of PMS stars in the near -- IR. In a companion paper (DJW03), we use this technique to study a sample of YSOs in the ρ Ophiuchi molecular cloud core. Using the Na interval at 2.2 µm, we can recover the effective temperatures of dwarf MK standards at a level below the uncer- tainty in the spectral type -- temperature conversion. The spectra also give us a good measure of the rotation velocity and continuum veiling. With the addition of a measurement of the relative flux between the (2 -- 0) 12CO bandhead and the Na interval, it is possible to deter- -- 20 -- mine the surface gravity of YSOs and to remove uncertainties in the temperature caused by a partial degeneracy with log g. The derived parameters are insensitive to extinction along the line -- of -- sight and need photometric information only to tie together the intensity scales in non -- contiguous high resolution spectra and to correct for differential reddening between 2.2 µm and 2.3 µm. It is clear from the results that the ability to work with weak features and to measure line shapes as well as equivalent widths can make high resolution near -- IR spectroscopy a valuable tool for studies of highly obscured stars and of young objects with strong excess infrared emission. The technique we develop here is robust enough to deal with sources with a range of different S/N ratios and with excess near -- IR emission. In the future, it will be worthwhile to investigate whether high resolution spectra of other near -- IR intervals could add useful information about the stellar parameters. We would also like to investigate more sophisticated matching routines that might make better use of the sensitivity of individual weak features to various stellar parameters. For studies of cooler, lower mass YSOs, it would be useful to produce models and line syntheses for lower temperature objects. We thank Chris Sneden for his advice and support in adapting MOOG for use with YSOs, Tom Greene for providing us with his data on MK standards, Kevin Luhman for making available his sample of Ophiuchus spectra, Ken Hinkle for help with our PHOENIX observations, and Russel White and Carlos Allende -- Prieto for helpful comments. Early phases of this work were supported in part by NSF grant AST 95-30695 to the University of Texas. A. Surface Gravity Diagnostics The ratio of photospheric equivalent widths from the Na doublet at 2.2 µm and the (2 -- 0) 12CO bandhead at 2.3 µm is sensitive to changes in surface gravity. Figure 5 shows that this ratio gives a good estimate of log g across the entire range in effective temperature and surface gravity relevant to the study of low mass PMS stars. In that figure, the photospheric equivalent widths in both the Na interval and the 12CO interval were calculated from model spectra synthesized from the NextGen stellar atmosphere models for log g = 3.5 -- 5.0 and Teff = 3000 -- 5000 K. The problem in studying YSOs is that they suffer both from significant extinction and reddening and often from the presence of excess continuum emission that can be larger -- 21 -- than the emission from the photospheres themselves. As a result, the measured equivalent width ratios do not accurately reflect the ratio of photospheric equivalent widths that would be relevant for comparison with Figure 5. What we would like to be able to do is to take the observed equivalent widths and, using as little additional data as possible and in a way as insensitive as possible to the effects of reddening and infrared excess, correct the observed equivalent width ratio to the photospheric equivalent width ratio. We outline here a procedure that uses near -- IR photometry to correct for differential reddening between 2.2 µm and 2.3 µm and low resolution spectroscopic data (if necessary) to correct for throughput differences between the high resolution spectra in the Na and 12CO intervals. In both cases, the corrections are usually quite small. To make it easier to use the equivalent width ratio as a diagnostic for surface gravity, the expressions below are geared to observable quantities. Starting with an absorption spectrum at high resolution, we define the measured equiv- alent width (MEW) in terms of the measured flux (Fλ) relative to the measured continuum (cλ): MEW = Z cλ − Fλ cλ dλ (A1) We define the photospheric equivalent width (PEW) to be the equivalent width of a photospheric absorption line without its value being altered by the presence of continuum veiling originating outside of the photosphere (rλ=(Fsource -- Fphot)/Fphot). We define, then, photospheric equivalent widths for the regions around the Na features at 2.2 µm (PEWNa) and the shortest wavelength part of the (2 -- 0) 12CO R -- branch at 2.3 µm (PEWCO). PEWNa = Z 2.212µm 2.202µm c2.2µm − F (λ) c2.2µm (1 + r2.2µm) dλ PEWCO = Z 2.3000µm 2.2925µm c2.3µm − F (λ) c2.3µm (1 + r2.3µm) dλ (A2) (A3) The photospheric continuum pλ is altered by infrared excess (increasing the continuum by a factor (1 + rλ)), probably due to thermal emission from a warm surrounding disk, and extinction (Aλ) along the line -- of -- sight. The corresponding expressions for the measured continuum (cλ) in the two wavelength regions of interest are: c2.2µm = p2.2µm(1 + r2.2µm)10−0.4A2.2µm erg s−1 cm−2 µm−1 (A4) -- 22 -- c2.3µm = p2.3µm(1 + r2.3µm)10−0.4A2.3µm erg s−1 cm−2 µm−1 (A5) Knowing the effective temperature (Teff) allows us to relate the photospheric continua to each other by assuming a blackbody dependence (B(T, λ)). χ(Teff) = p2.3µm p2.2µm = Bλ(Teff, 2.3µm) Bλ(Teff, 2.2µm) (A6) By taking the ratio of equations A4 & A5 and substitution of equation A6, we arrive at an expression for a quantity we can measure with low resolution spectra or with cross dispersed high resolution spectra covering the full range of the two wavelength intervals: the ratio of measured continua at two different wavelengths. c2.2µm c2.3µm = (1 + r2.2µm)10−0.4A2.2µm χ(Teff)(1 + r2.3µm)10−0.4A2.3µm (A7) Re -- arranging equation A7 and substituting into the ratio of equations A2 & A3 removes the dependence on any continuum veiling: PEWNa PEWCO = MEWNa MEWCO c2.2µm c2.3µm χ(Teff)10−0.4(A2.3µm−A2.2µm) (A8) This expression provides terms on the right hand side that we can measure with spec- troscopy and photometry allowing us to utilize the surface gravity dependence on photo- spheric equivalent ratios in models as illustrated in Figure 5. REFERENCES Ali, B., Carr, J.S., DePoy, D.L., Frogel, J.A., & Sellgren, K. 1995, AJ, 110, 2415 Allen, C. W. 2000, Allen's Astrophyical Quantities (New York, Springer -- Verlag) Alonso, A., Arribas, S., & Martinez -- Roger, C. 1996, A&A, 313, 873 Baldwin, J.R., Frogel, J.A., & Persson, S.E. 1973, ApJ, 184, 427 Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 1998, A&A, 337, 403 Barsony, M., Kenyon, S.J., Lada, E.A., & Teuben P.J. 1997, ApJS, 112, 109 -- 23 -- Barsony, M., Koresko, C., & Matthews, K. 2003, ApJ, 591, 1064 Bontemps, S. et al. 2001, A&A, 372, 173 Casali, M. M. & Matthews, H. E. 1992, MNRAS, 258, 399 Casali, M. M. & Eiroa, C. 1996, A&A, 306, 427 Comer´on, F., Rieke, G. H., Burrows, A., & Rieke, M. J. 1993, ApJ, 416, 185 D'Antona, F. & Mazzitelli, I. 1997, Mem. Soc. Astron. Italiana, 68, 823 Delfosse, X., Forveille, T., Perrier, C., & Mayor, M. 1998, A&A, 331, 581 de Jager, C. & Nieuwenhuijzen, H. 1987, A&A, 177, 217 de Medeiros, J. R. & Mayor, M. 1999, VizieR Online Data Catalog, 413, 990433 Doppmann, G.W., Jaffe, D.T., White, R.J. 2003 AJ, submitted (DJW03) Duquennoy, A. & Mayor, M. 1988, A&A, 200, 135 Duquennoy, A. & Mayor, M. 1991, A&A, 248, 485 Edvardsson, B., Andersen, J., Gustafsson, B., Lambert, D. L., Nissen, P. E., & Tomkin, J. 1993, A&A, 275, 101 Fekel, F.C. 1997, PASP, 109, 514 Ghez, A. M., Neugebauer, G., & Matthews, K. 1993, AJ, 106, 2005 Glebocki, R. & Stawikowski, A. 2000, Acta Astronomica, 50, 509 Goorvitch, D. & Chackerian, C. Jr. 1994, ApJS, 91, 483 Gray, D. F. 1992, The Observation and Analysis of Stellar Photospheres, (New York, Cam- bridge Univ. Press) Gray, R. O., Graham, P. W., & Hoyt, S. R. 2001, AJ, 121, 2159 Greene, T.P., & Young, E.T. 1992, ApJ, 395, 516 Greene, T. P., Tokunaga, A. T., Toomey, D. W., & Carr, J. B. 1993, Proc. SPIE, 1946, 313 Greene, T. P., Wilking, B. A., Andr´e, P., Young, E. T., & Lada, C. J. 1994, ApJ, 434, 614 Greene, T. P. & Meyer, M. R. 1995, ApJ, 450, 233 -- 24 -- Greene, T. P. & Lada, C. J. 1996, AJ, 112, 2184 Greene, T. P. & Lada, C. J. 1997, ApJ, 114, 2157 Greene, T. P. & Lada, C. J. 2000, ApJ, 120, 430 Greene, T. P. & Lada, C. J. 2002, AJ, 124, 2185 Gullbring, E., Calvet, N., Muzerolle, J., & Hartmann, L. 2000, ApJ, 544, 927 Hartmann, L., Stauffer, J. R., Kenyon, S. J., & Jones, B. F. 1991, AJ, 101, 1050 Hauschildt P.H., Allard, F. & Baron, E. 1999, ApJ, 512, 377 Hearnshaw, J. B. 1974, A&A, 34, 263 Hearnshaw, J. B. 1974, A&A, 36, 191 Hinkle, K. H., Cuberly, R. W., Gaughan, N. A., Heynssens, J. B., Joyce, R. R., Ridgway, S. T., Schmitt, P., & Simmons, J. E. 1998, Proc. SPIE, 3354, 810 Huang, S. 1961, ApJ, 134, 12 Johns -- Krull, C. M. & Valenti, J. A. 2001, ApJ, 561, 1060 Keenan, P. C. & McNeil, R. C. 1989, ApJS, 71, 245 Kenyon, S. J. & Hartmann, L. 1995, ApJS, 101, 117 Kenyon, S. J., Brown, D. I., Tout, C. A., & Berlind, P. 1998, ApJ, 115, 2491 Kirkpatrick, J. D., Henry, T. J., & McCarthy, D. W. 1991, ApJS, 77, 417 Kirkpatrick, J.D., Kelly, D.M., Rieke, G.H., Liebert, J. Allard, F., & Wehrse, R. 1993, ApJ, 402, 643 Kleinmann, S. G., & Hall, D. N. B. 1986, ApJS, 62, 501 Kurucz, R. L. 1994, Atomic Data for Opacity Calculations, Kurucz CD -- ROM No. 1 Lada, C. J. 1987, IAU Symp. 115: Star Forming Regions, 115, 1 Lancon, A., & Rocca -- Volmerange, B. 1992, A&AS, 96, 593 Leggett, S.K., Allard, F., Berriman, G., Dahn, C.C., and Hauschildt, P.H. 1996, ApJS, 104, 117 -- 25 -- Livingston, W. & Wallace L. 1991 An Atlas of the Solar Spectrum in the Infrared from 1850 to 9000 cm-1 (1.1 to 5.4 microns), N.S.O. Technical Report #91-001, July 1991 Luhman. K. L. & Rieke, G. H. 1999, ApJ, 525, 440 Marsakov, V. A. & Shevelev, Y. G. 1988, Bulletin d'Information du Centre de Donnees Stellaires, 35, 129 McCaughrean, M. J. 2001, IAU Symposium, 200, 169 McLean, I. S., Becklin, E. E., Figer, D. F., Larson, S., Liu, T., & Graham, J. 1995, Proc. SPIE, 2475, 350 McWilliam, A. 1990, ApJS, 74, 1075 Merrill, K.M., & Ridgway, S.T. 1979, ARA&A, 17, 9 Meyer, M. R., Edwards, S., Hinkle, K. H., & Strom, S. E. 1998, ApJ, 508, 397 Mountain, C.M., Robertson, D.J., Lee, T.J., & Wade, R. 1990, in Instrumentation in Astronomy, Proc. SPIE, p. 25 Padgett, D. L. 1996, ApJ, 471, 847 Palla, F. & Stahler, S. W. 1999, ApJ, 525, 772 Palla, F. & Stahler, S. W. 2000, ApJ, 540, 255 Ram´irez, S.V., DePoy, D.L., Frogel, J.A., Sellgren, K., & Blum, R.D. 1997, AJ, 113, 1411 Siess, L., Dufour, E., & Forestini, M. 2000, A&A, 358, 593 Simon, M. et al. 1995, ApJ, 443, 625 Simon, M., Dutrey, A., & Guilloteau, S. 2000, ApJ, 545, 1034 Sneden, C. 1973, Ph.D. Thesis, University of Texas at Austin Stahler, S. W. 1988, PASP, 100, 1474 Strassmeier, K. W. A., Granzer, T., Scheck, M., & Weber, M. 2000, A&AS, 142, 275 Strom, K. M., Kepner, J., & Strom, S. E. 1995, ApJ, 438, 813 Taylor, B. J. 1995, PASP, 107, 734 -- 26 -- Unsold, A. 1955, Physik der Sternatmospharen (2nd ed.; Berlin: Springer -- Verlag) Vogt, S. S., Penrod, G. D., & Soderblom, D. R. 1983, ApJ, 269, 250 Wallace, L., & Hinkle, K. 1996, ApJS, 107, 312 Webb, R. A., Zuckerman, B., Platais, I., Patience, J., White, R. J., Schwartz, M. J., & McCarthy, C. 1999, ApJ, 512, L63 White, R. J., Ghez, A. M., Reid, I. N., & Schultz, G. 1999, ApJ, 520, 811 White, R. J. & Ghez, A. M. 2001, ApJ, 556, 265 Wilking, B. A., & Lada, C. J. 1983, ApJ, 274, 698 Wilking, B. A., Lada, C. J., & Young, E. T. 1989, ApJ, 340, 823 Wright, G. S., Mountain, C. M., Bridger, A., Daly, P. N., Griffin, J. L., & Ramsay Howat, S. K. 1993, Proc. SPIE, 1946, 547 This preprint was prepared with the AAS LATEX macros v5.0. -- 27 -- Ca I Fe I Sc I Sc I Si I Si I Na I Fe I Na I Ca I V I R Branch 58 43 60 41 64 37 66 35 62 39 Ti I Fig. 1. -- The intervals within the 2.0 -- 2.4 µm K window used for spectral synthesis analysis of high resolution data. The spectra are from a synthesis of a Teff = 4000 K, log g = 3.5 NextGen atmosphere model (Hauschildt et al. 1999), assuming a resolving power λ/∆λ = 50, 000. The intensity is normalized to the photospheric continuum. The Na interval shows the numerous neutral atomic species present in the photospheres of cool stars (top). The 12CO interval shows the bandhead and the v=2 -- 0 R -- branch transitions of 12CO (bottom). The vertical lines show the boundaries for the intervals used to compute the equivalent width ratios plotted in Figure 5. -- 28 -- Fig. 2. -- A grid of spectral syntheses based on NextGen (Hauschildt et al. 1999) atmosphere models. These spectra illustrate the change in Na line shape and depth and the variations in Si/Sc for temperatures from 3300 K to 4200 K for log g fixed at 4.0. Spectra have been smoothed to a R = 50,000. -- 29 -- 1.2 1 0.8 0.6 0.4 1.2 1 0.8 0.6 0.4 1.2 1 0.8 0.6 0.4 RMS=0.0395 RMS=0.0321 RMS=0.0362 2.202 2.204 2.206 2.208 2.21 2.212 Fig. 3. -- Fits to a noisy artificial spectrum of the Na interval. The thin solid line shows an artificial spectrum for Teff = 3600 K, log g = 4.0, v sin i = 15 km s−1, and rNa = 1.0. We have smoothed this spectrum to R = 50,000 and added Gaussian random noise to the spectrum to produce a S/N = 30 (per pixel with R = 240,000 pixels). Overlaid on each of the 3 panels is a noiseless synthetic spectrum (bold lines) for log g = 4.0, Teff = 4000K (top), Teff = 3600K (middle), and Teff = 3200K (bottom). In each panel, the best fit was found holding Teff fixed at the listed value but allowing rNa and v sin i to vary. The difference between the noisy artificial and noiseless spectra is displayed at the bottom of each panel. The vertical dashed lines are the subintervals chosen in this case to enclose regions with significant line flux over which we will compute the RMS difference between the noisy artificial spectrum and the grid of noiseless synthetic spectra. In the lower left of each panel is the RMS difference between the noisy and noiseless spectrum normalized to the number of points in the two subintervals that enclose the photospheric lines. -- 30 -- ) 1 - s m k ( n o i t a t o R 30 20 10 3.0 1.0 0.3 ) a N r ( g n i l i e V ) 2 - s m c ( g g o L 4.8 4.4 4.0 3.6 3200 3600 Teff (K) 4000 Fig. 4. -- Variations of the RMS error in the Teff -- v sin i, Teff -- rNa, and Teff -- log g planes for a comparision of the noisy artificial spectrum from Figure 3 (where fluxes are normalized to one) to noiseless synthetic spectra. The contours show the RMS deviation of the noisy spectrum from the model at each point in the parameter space. We produced each two -- parameter plot while holding the other two variables fixed at the values matching the correct values for the target spectrum. The error bar at the RMS minimum in each plot represents the standard deviation of the best fit value for S/N = 30 spectra with different noise seeds. -- 31 -- Log g 3.5 Log g 4.0 Log g 4.5 Log g 5.0 0.6 0.4 0.2 0 -0.2 3000 3500 4000 4500 5000 5500 Fig. 5. -- The ratio of CO interval equivalent width to Na interval equivalent width as a function of temperature, plotted for surfaces gravities between log g = 3.5 and log g = 5.0. Equivalent widths were computed over the Na and 12CO intervals as defined in Figure 1. The relatively flat shape of these isogravity lines with temperature illustrates the value of the 12CO/Na line flux ratio as a good diagnostic of surface gravity. -- 32 -- 2.204 2.206 2.208 2.21 2.212 3.5 Log g 4.0 4.5 Na 2.5 2 1.5 1 0.5 0 2.202 0.2 0.1 0 -0.1 -0.2 3400 3600 3800 4000 1.5 1 0.5 2.292 2.294 2.296 2.298 Fig. 6. -- Demonstration of simultaneous derivation of Teff and log g. The top panel shows a S/N = 50 synthetic spectrum for (Teff , log g)= (3800K, 4.0). Below that we show differences between this spectrum and models for (Teff, log g) = (3500K, 3.5), (3800K, 4.0), and (4000K, 4.5) illustrating the effects of the broad minimum in the errors as Teff and log g increase simultaneously. The middle panel shows the difference between the equivalent width of the Na and CO interval for these (Teff, log g) pairs and (Teff, log g) = (3800K, 4.0), normalized to the equivalent width of the (3800K, 4.0) spectrum. The bottom panel shows the associated strong variations in the spectra of the CO interval. -- 33 -- Fig. 7. -- Standard deviation of the derived value of Teff (derived using the Na interval only) as a function of S/N. To construct this plot, we used a noisy artificial spectrum with Teff=4200K, log g=4.0, v sin i=15 kms−1, and rNa=1.0 as the target and fit for the best temperature allowing all parameters except log g to vary. We compared the target and model spectra at sampling points spaced every ∆λ/λ=(1/240,000) along spectra smoothed to a resolving power of 50,000. The S/N is that appropriate to data binned to channels ∆λ/λ=(1/50,000) wide. -- 34 -- log g = 4.5 log g = 4.0 log g = 3.5 0.04 0.02 0 -0.02 -0.04 3500 4000 4500 5000 Fig. 8. -- The effect on the best -- fit value of Teff of assuming an incorrect value for log g. At each temperature, a synthetic target spectrum with log g=4.0 and two grids of MOOG models with log g=3.5 and log g=4.5 were created. We then fit for the temperature of the target spectrum using each of the grids with the different values of log g. The solid lines show the difference between the logarithm of the best -- fit Teff to the target spectrum derived by using the log g=3.5 (bottom) or 4.5 (top) grids and the logarithm of the target spectrum Teff. -- 35 -- B-V colors Fig. 9. -- Test of our effective temperature determinations (vertical axis) against temperature determinations available in the literature for MK standards (horizontal axis). The small vertical error bars represent ±1σ uncertainties due to the noise in the observed spectra, determined by calculations similar to those used in Figure 7 at each temperature and at a S/N comparable to that of the data. The horizontal error bar in the lower right corner of the Figure shows the 1 σ uncertainty in the spectral type to Teff conversion at 4000 K (de Jager & Nieuwenhuijzen 1987). We placed the symbols in the x direction by converting the spectral types (Table 1) to Teff using various spectral type -- Teff relations: open circles (Alonso et al. 1996), open triangles (Ali et al. 1995), open squares (Allen 2000) vertical error bars (de Jager & Nieuwenhuijzen 1987). Unless otherwise marked, all sources are dwarfs, luminosity class V. The open pentagons show the Teff derived from the reported B -- V colors using the conversion relation of Kenyon & Hartmann (1995). The dashed line shows the relation Teff(near -- IR) = Teff(optical). -- 36 -- Table 1. Observed MK standards Standard Star Type Spectral Optical Near -- IR Optical Near -- IR Metal- licity [Fe/H] v sin i (km s−1) v sin i (km s−1) c Teff (K) d Teff (K) HD 117176a HR 4496b HR 4496a HD 185144a G4V1 G8V1 G8V1 K0V1 K2V1 GL 28b K4V1 HR 5568b M0V2 GL 338Ab GL 338Aa M0V2 HD 131976a M1.5V1 M2V1 M4V2 HD 188512a G8IV1 HD 142091a K1IVa1 GL 15Ab GL 402b 5636 5439 5439 5152 4838 4539 3837 3837 3589 3523 3289 51005 48005 5980 5400 5460 5260 4880 4580 3660 3650 3610 3630 3290 4700 4680 108 <158 <158 <158 2.59 <128 2.910 2.910 1.411 <2.910 <2.310 1.813 1.914 <3 <9 <3 <3 <9 <9 <9 <3 10 <9 <9 <3 <3 −0.113 −0.144 −0.144 −0.233 −0.056 0.0167 ... ... ... ... ... −0.305 −0.045 S/N 120 300 110 120 180 170 155 110 75 150 110 130 100 aObservations made using PHOENIX on the KPNO 4 -- meter bObservations made using NIRSPEC on Keck cUsing Teff -- spectral type relation of de Jager & Nieuwenhuijzen (1987) for luminosity class V sources unless otherwise noted dAssumed log g=4.5 for dwarfs and log g=3.5 for the two sub -- giants sources (McWilliam 1990) References. -- (1) Keenan & McNeil (1989) (2) Kirkpatrick et al. (1991) (3) Hearnshaw (1974b) (4) Hearnshaw (1974a) (5) McWilliam (1990) (6) Marsakov & Shevelev (1988) (7) Taylor (1995) (8)Glebocki & Stawikowski (2000) (9) Strassmeier et al. (2000) (10) Delfosse et al. (1998) (11) Duquennoy & Mayor (1988) (12) Vogt et al. (1983) (13) de Medeiros & Mayor (1999) (14) Fekel (1997) -- 37 -- Table 2. Recovered parameters with rotation (v sin i = 25 km s−1) and veiling (rNa = 2.0) added to MK standards MK Best Fit Recovered Recovered Recovered RMS Standard HD 117176 HR 4496 HD 185144 GL 338A HD 131976 HD 188512 HD 142091 Teff (K) 5980 5400 5240 3650 3610 4700 4680 Teff (K) 5240 5230 5000 3765 3580 4620 4620 v sin i (km s−1) Veiling (rNa) 27.0 29.0 29.0 25.0 30.0 30.0 29.0 3.8 2.6 2.7 3.0 2.5 2.3 2.5 0.00271 0.00152 0.00189 0.00333 0.00244 0.00165 0.00209
astro-ph/0104325
1
0104
2001-04-19T17:27:05
Sco X-1: Energy Transfer from the Core to the Radio Lobes
[ "astro-ph" ]
The evolution of the radio emission from Sco X-1 is determined from a 56-hour continuous VLBI observation and from shorter observations over a four-year period. The radio source consists of a variable core near the binary, and two variable compact radio lobes which form near the core, move diametrically outward, then fade away. Subsequently, a new lobe-pair form near the core and the behavior repeats. The differences in the radio properties of the two lobes are consistent with the delay and Doppler-boosting associated with an average space velocity of 0.45c at 44 deg to the line of sight. Four lobe speeds, between 0.32c and 0.57c, were measured for several lobe-pairs on different days. The speed during each epoch remained constant over many hours. The direction of motion of the lobes over all epochs remained constant to a few degrees. Two core flares are contemporaneous with two lobe flares after removal of the delay associated with an energy burst moving with speed >0.95c in a twin-beam from the core to each lobe. This is the first direct measurement of the speed of energy flow within an astrophysical jet. The similarity of the core and lobe flares suggests that the twin-beam flow is symmetric and that the core is located near the base of the beam.
astro-ph
astro-ph
Sco X-1: Energy Transfer from the Core to the Radio Lobes National Radio Astronomy Observatory, Charlottesville, VA 22903 E. B. Fomalont [email protected] B. J. Geldzahler & C. F. Bradshaw School of Computational Sciences, George Mason University, Fairfax, VA 22030 [email protected] & [email protected] ABSTRACT The evolution of the radio emission from Sco X-1 is determined from a 56-hour continuous VLBI observation and from shorter observations over a four-year period. The radio source consists of a variable core near the binary, and two variable compact radio lobes which form near the core, move diametrically outward, then fade away. Subsequently, a new lobe-pair form near the core and the behavior repeats. The differences in the radio properties of the two lobes are consistent with the delay and Doppler-boosting associated with an average space velocity of 0.45c at 44◦ to the line of sight. Four lobe speeds, between 0.32c to 0.57c, were measured for several lobe-pairs on different days. The speed during each epoch remained constant over many hours. The direction of motion of the lobes over all epochs remained constant to a few degrees. Two core flares are contemporaneous with two lobe flares after removal of the delay associated with an energy burst moving with speed βj > 0.95 in a twin-beam from the core to each lobe. This is the first direct measurement of the speed of energy flow within an astrophysical jet. The similarity of the core and lobe flares suggests that the twin-beam flow is symmetric and that the core is located near the base of the beam. Subject stars:individual(Sco X-1) -- X-rays:stars headings: binaries:close -- galaxies:jets -- radiocontinuum: stars -- stars:neutron -- 1. Introduction its X-ray color-color diagram. Sco X-1 is a low-mass binary system (LMXB) with strong, persistent X-ray emission. It is a 'Z- type' LMXB, named for a characteristic shape of It also ex- hibits quasi-periodic X-ray oscillations (Hasinger & van der Klis 1989; van der Klis et al. 1996). The binary has an orbital period of 0.787d (Gottlieb et al. 1975), the degenerate object is probably a neutron star, with a one solar mass companion of unknown spectral type (Cowley & Crampton 1975). From eight Very Long Baseline Array (VLBA) observations between 1995 and 1998, a distance of 2.8 ± 0.3 kpc was obtained from measure- ments of the trigonometric parallax of Sco X- 1 (Bradshaw et al. 1999). The radio emission also varied over tens of minutes and relativis- tic motion was detected. Thus, Sco X-1 ex- hibited the properties of Galactic-jet sources (Mirabel & Rodriguez 1999) with similarities to extragalactic radio sources (Blandford & Rees 1974). However, these observations were too short and too intermittent to determine the properties of Sco X-1. We, therefore, observed Sco X-1 with milliarc- second (mas) resolution over a 56-hour period on 1999 June 11-13 (MJD 51340-2). Extensive opti- cal and X-ray observations were also made. This letter reports on the general properties of Sco X- 1 1 with emphasis on the kinetic properties of the source. Further discussions of the radio observa- tions and the nature of Sco X-1 are given in Paper II (Fomalont et al. 2001). The nature of the x-ray emission and accretion processes in Sco X-1, and the interaction of the radio core and the binary system are given elsewhere (Titarchuk et al. 2001; Bradshaw et al. 2002; Geldzahler et al. 2002). 2. The Radio Observations: The observations in 1999 June consisted of seven consecutive 8-hour VLBI observations cy- cling among three different arrays: the VLBA with the VLA, the APT (Asia-Pacific Telescope: Aus- tralia Telescope Compact Array, Ceduna, Harte- beesthoek, Kashima, Mopra, Parkes and Shang- hai), and the EVN (European VLBI Network: Ef- felsberg, Jodrell Bank, Medicina, Noto, Wester- bork, plus Hartebeesthoek and the Green Bank 140-ft). With the VLBA, 1.7 GHz and 5.0 GHz ob- servations were interleaved in time. With the APT and EVN, observations were made at 5.0 GHz only. All data associated with the observations of Sco X-1 were calibrated using simultaneous ob- servations of a nearby background 10-mJy radio source1, by which all images were registered to an accuracy of < 0.1 mas (Bradshaw et al. 1999). Because of the significant flux density variations and component motions, images were made from data sets as short as 50 minutes (snap-shots) in order to follow the changes in the radio emission. 3. The Radio Snap-shots of Sco X-1 The sequence of the radio snap-shots of Sco X-1 during the 1999 June observations is displayed in Figure 1. More frequent, higher-resolution 5-GHz snap-shots are displayed in Paper II. The radio structure showed three components: a radio core within a few mas of the binary; a compact moving component north-east of the core, called the NE lobe, and a fainter component south-west of the core, called the SW lobe, detected most strongly 1This source is one of the two radio objects which were previ- ously thought to be associated with Sco X-1. They are un- related background objects (Fomalont & Geldzahler 1991) at a distance from Sco X-1 of 70′′ in position angle 30◦. The radio emission from Sco X-1 extends < 0.1′′ in posi- tion angle 54◦. on MJD 51342. Occasionally, faint extended emis- sion was detected between the core and NE lobe. The flux density variations of the three ma- jor components are shown in Figure 2. Sco X-1 evolved through several stages during these obser- vations. The first NE lobe disappeared at MJD 51340.4 after reaching a distance of 20 mas from the core. The core component, on the other hand, increased in flux density during this period. For the next nine hours, Sco X-1 consisted of only the core component. At MJD 51340.7 the core bright- ened and by MJD 51340.9 a new NE lobe (#2) emerged from the core. Some extended emission between these two separating components per- sisted until MJD 51341.3, during which time the flux density of both components decreased. Over the next 30 hours the NE lobe moved directly away from the core, reaching a separation of 50 mas. Both the core and NE lobe varied significantly in flux density; The NE lobe peak (N3) at MJD 51341.1 and the core peak (C4) at MJD 51341.9 reached 20 mJy at 5 GHz. There was no appar- ent correlation of the core flares with the phase of the binary orbit. The SW lobe, detected most strongly on MJD 51342, was located about half as far from the core as the NE lobe. For the observations between 1995 and 1998, Sco X-1 was relatively weak and an accurate motion of the NE lobe could not be measured. The one exception was the observations on 1998 February 27-28 1998 (MJD 50871-2), consisting of two six-hour VLBA observations separated by 18 hours, and these images are shown in Paper II, Figure 5. On MJD 50871, the NE lobe was about 20 mas from the core, and moved away at a speed similar to that in 1999 June. The SW lobe was also present and its motion, in the opposite direction as the NE lobe, was measured. 4. Lobe Velocities and Relativistic Effects Figure 3 shows the observed separation of the NE and SW lobes from the core. For the 1999 June observations the velocity in the plane of the sky of the NE lobe (#1) was v = 0.73 ± 0.07 mas hr−1. (For a measured distance to Sco X-1 of 2.8 kpc, one mas = 2.8 AU = 4.19 × 108 km. A projected velocity of 1 mas hr−1 = 0.388c.) After emerging from the core, the NE lobe (#2) lobe speed was v = 1.74 ± 0.16 mas hr−1. At MJD 51341.4 the 2 NE lobe flared and its speed abruptly decreased to v = 1.25 ± 0.05 mas hr−1. Although the NE lobe varied considerably in flux density over the next 30 hours, its speed remained nearly constant. For the 1998 February observations, the velocity of the NE lobe on MJD 50871 was v = 1.11 ± 0.06 mas hr−1, with no significant departure from uniform motion. The speeds for the weaker SW lobes were deter- mined less accurately. On MJD 50872 v = 0.6±0.2 mas hr−1, and on MJD 51342 v = 0.5 ± 0.3 mas hr−1. Further evidence of the motion of the SW lobe is described in Paper II which shows that whenever the SW lobe is detected, its distance from the core is always about half that of the NE lobe. Since the NE lobe was clearly moving away from the core, this ratio is also equal the ratio of the speed of the SW lobe, βSW , to the speed of the NE lobe, βN E and is βSW /βN E = 0.51 ± 0.02 us- ing all observations of Sco X-1. Such lobe velocity differences have been observed in other Galactic- jet sources (Mirabel & Rodriguez 1999) and are caused by delay effects from their relativistic mo- tion. The expected Doppler-boosting (attenuation) of the approaching (receding) component was ob- served in Sco X-1. An analysis in Paper II derives an average flux density ratio between the SW and NE lobes of 0.12 ± 0.03. On the other hand the spectral index, α (S∝ να), of the two lobes should not be affected significantly by a large space mo- tion and, indeed, both lobes had α ≈ −0.6 dur- ing the flaring states. The lobe angular diame- ters were also similar. This comparison of the ra- dio emission from the NE and SW lobes suggests that they are intrinsically similar, but their ap- pearances are strongly affected by the delay and Doppler effects of their space motion. With an accurate distance to Sco X-1, the lobe space motion can be determined solely from the observed lobe kinetics. Although four speeds be- tween 0.28c and 0.68c in the plane of the sky were measured for the NE lobe, there were two periods when the lobe speed remained constant for over one day: MJD 50871-2 with βN E = 0.43c ± 0.02c and MJD 51341-2 with βN E = 0.49c ± 0.02c. Thus, we adopt a typical speed of the NE lobe of βN E = 0.46c ± 0.08c with the estimated error derived from the spread of the measured speeds and the distance uncertainty to Sco X-1. Since the position angle of the radio axis of Sco X-1 in the plane of the sky over the five year observation period remained within a few degrees of 54◦ (Pa- per II), the direction of the space velocity of the lobes was also constant to this same level. Hence, any change in the observed speed of the lobes is caused by changes in speed and not in direction. With the derived values of βN E and βSW /βN E, the average space velocity of the lobes can be de- termined (Blandford et al. 1977): βSW βN E = 0.51 ± 0.02 = βN E = 0.46 ± 0.08 = 1 − β cos(θ) 1 + β cos(θ) β sin(θ) 1 − β cos(θ) (1) (2) where β is the true speed of the lobes moving in direction θ with respect to the observer. The average space velocity of the lobes is then β = 0.45 ± 0.03; θ = 44◦ ± 6◦. The quoted errors are the one-sigma uncertainty. The four measured space velocities for different lobe-pairs were: 0.32c, 0.43c, 0.46c and 0.57c. The flux density ratio for the SW to NE lobe from Doppler-boosting expected from the derived space velocity is R = (βSW /βN E)k−α (Blandford & Konigl 1979). For k=2, which is appropriate for an optically thin discrete com- ponent, and α = −0.6 as measured, we predict R= (0.51)−2.6 ≈ 0.17. However, if the mag- netic field distribution in the lobes is anisotropic, the flux ratio could be considerably smaller (Cawthorne 1991). The observed flux density ra- tio, R=0.12 is in reasonable agreement with that expected from the Doppler boosting which was derived solely from the lobe motions. 5. The Correlation of Flux Density Varia- tions and the Jet Speed The large flux density variations of the core and the lobes during the 1999 June observations are shown in Figure 2. The flare characteristics of all components were similar; a factor of 2 to 10 in- crease in flux density; −0.3 < α < −0.6 (except for flare C1); a sharper rise time than decay time; a flare width of about three hours; and a lobe an- gular size < 3 mas. In Paper II we suggest that the NE and SW lobes are moving hot-spots whose internal and ki- netic energies are supplied by the disruption of the 3 energy flow in an oppositely-directed twin-beam formed near an accretion disk around the neu- tron star. Therefore, we searched for a correlation of the intensity variations of the lobes versus the core using the following model: (1) A core flare re- sponds in an unspecified manner to an event near the binary system; (2) This event is also associ- ated with an energy surge which propagates down the twin-beam with an average velocity of βj; (3) Each lobe flares when this increased energy flux reaches them. With this model the expected de- lay, as seen by an observer, between a core flare and its manifestation in the NE and SW lobes can be determined since the lobe space velocity, β and θ, are accurately known. The expected delays, τN E, τSW , are: τN E = (t1 − t0)β τSW = (t1 − t0)β (1 − βjcosθ) (βj − β) (1 + βjcosθ) (βj − β) (3) (4) where t0 is the time of ejection of the lobes from the core, and t1 is the time of a core flare. The (βj −β)−1 factor is related to the time for the flare 'information' to travel in the beam from the core to the lobes. Figure 4 shows the flux density variations of the three components, after removal of the delay as- sociated with the NE and SW lobes for βj = 1.0 and near the flare peaks for βj = 0.9. The tempo- ral correlations for the flares C3-N3 (τN E = 2.8h), C4-N4 (τN E = 7.4h) and C3-S3 (τSW = 19.7h) with βj = 1.0 are excellent, and the sensitivity of these correlations to βj are illustrated from those points plotted for βj = 0.9. A fit of the correlation of the three observed flare-pairs as a function of the jet speed gives a formal solution, βj = 1.02 ± 0.04 (Paper II). Even with these lim- ited statistics and an over-simplified model, these observations suggest that the speed of energy flow in the beam is > 0.95c. This result is the first di- rect measurement of energy flow velocity in an as- trophysical beam. The velocity close to the speed of light suggests that the energy flux is probably carried by leptons with γ > 3, rather than with protons or heavier particles moving at velocities considerably less than c (Blandford & Rees 1974). For flare 3, the Doppler-corrected flux densities, the general shapes, and spectral properties of the 4 NE and SW lobes are similar. This suggests that the energy flow in the twin-beam is symmetric, and that the production of radio energy in each lobe depends mainly on the beam energy flow, rather than on interaction details at the working surfaces in the lobes. For flare 4 the flux density of the core is about ten greater than that for the NE lobe. (This flare in the SW lobe did not oc- cur until thirty hours after the termination of the experiment.) Possible explanations on the relative weakness of the NE lobe for this flare are discussed in Paper II. The similar properties of the two core flares with the lobe flares are surprising since the pro- cesses producing the radio emission in the lobes and in the core are thought to be different. Since these core flares are clearly associated with the energy flow in the beam, we suggest that the core component is located near the base of the beam rather than within the binary system. A symmet- ric 'core' component at the base of the beam to the south-west of the binary should also be present, but would be strongly Doppler-attenuated. Fur- ther discussion of the relationship between the ra- dio core and the binary system are given elsewhere (Geldzahler et al. 2002), and includes a discussion of flare 2 which is associated with the formation of a new pair of lobes, and flare 1 which is anti- correlated with NE lobe in the sense that the NE lobe vanishes when the core reached a maximum. 6. Conclusions The relatively simple and recurrent nature of the radio properties of Sco X-1 permit the deter- mination of its kinetics. The compact lobes move directly away from the core with an average ve- locity of 0.45c ± 0.03c at an angle 44◦ ± 6◦ to the line of sight, with a range for four different lobe- pairs between 0.32c to 0.57c. The lobe advance re- mains constant for many hours despite significant flux density variations. The excellent correlations among some of the core and lobe flares suggest that energy flux within the twin-beam flows at a speed > 0.95c, that the energy flow is symmetric in the twin-beams, that the radio luminosity of the lobes and core depends strongly on the energy flux in the beam but weakly on other environmen- tal parameters, and that the location of the radio core is more likely to be at the base of the beam rather than near the binary system or accretion disk. Hasinger, G. & van der Klis, M. 1989, A&A, 225, 79 Mirabel, I. F. & Rodriguez, L. F. 1999, A&A Rev., 37, 409 Titarchuk, L. G., Bradshaw, C. F., Geldzahler, B. J. & Fomalont, E. B. 2001, in preparation. van der Klis, M., Swank, J. H., Zhang, W., Ja- hoda, K., Morgan, E. H., Lewin, W. H. G., Vaughan, B. & van Paradijs, J. 1996, ApJ, 469. L1 The National Radio Astronomy Observatory is a facility of the National Science Foundation, op- erated under cooperative agreement by Associated Universities, Inc. We thank the European VLBI Network (EVN) and the Asia-Pacific Telescope (APT) for their support and observation time. The data were correlated with the VLBA correla- tor in Socorro and the Penticton Correlator which is supported by the Canadian Space Agency. It is a pleasure to thank Dr. Jean Swank for grant- ing us RXTE time, to Dr. Tasso Tzioumis for help with the APT scheduling, to Dr. Sean Dougherty for processing the APT data. We also thank Dr. Michael McCollough for comments on the draft. REFERENCES Blandford, R. D. & Konigl, A. 1979, ApJ, 232, 34 Blandford, R. D. , McKee, C. F. & Rees, M. J. 1977, Nature, 267, 211 Blandford, R. D. & Rees, M. J. 1974, MNRAS, 169, 395 Bradshaw, C. F., Fomalont, E. B. & Geldzahler, B. J. 1999, ApJ, 512, L121 Bradshaw, C. F., Geldzahler, B. J & Fomalont, E. B. 2002, in preparation Cawthorne, T. Jets A. and P. Univ. Press), 187 in V. Beams ed. Hughes(Cambridge:Cambridge 1991, Astrophysics, in Cowley, A. P. & Crampton, D. 1975, ApJ, 201, L65 Fomalont. E. B. & Geldzahler, B. J. 1991, ApJ, 383, 289 Fomalont, E. B., Geldzahler, B. J. & Bradshaw, C. J. 2001, ApJ, submitted [Paper II] Geldzahler, B. J., Fomalont, E. B. & Bradshaw C. J. 2002, in preparation Gottlieb, E. W., Wright, E. L. & Liller, W. 1975, ApJ, 195, L33 This 2-column preprint was prepared with the AAS LATEX macros v5.0. 5 Fig. 1. -- Snap-shots of Sco X-1 with 10 × 5 mas resolution during the 1999 June observations. The 5.0 GHz and 1.7 GHz snapshots are shown side by side with the time axis on the left of the two vertical blocks. Each image has been rotated 36◦ with respect to North, indicated by the arrow near the center of the figure. The contour levels at 5.0 GHz are 0.5 mJy × (−1, 1, 2, 4, 8 . . .), and at 1.7 GHz are 0.7 mJy × (−1, 1, 2, 4, 8 . . .). The tick marks are separated by 10 mas along the abscissa and 5 mas along the ordinate. Snapshots labeled with A are from the APT observations, those with E are from the EVN observations. All others snapshots are from the VLBA. The vertical dashed-line indicates the location of the core (determined from previous radio observations), the line to the left the location of the NE lobe (#1) and the NE lobe (#2). The line to the right, for MJD 51342 only, shows the location of the SW lobe. 6 Fig. 2. -- The Flux Density of the Components in Sco X-1 during the 1999 June Observations. The 5-GHz flux density for the core and NE lobe, and the 1.7-GHz for the SW lobe are plotted. (The weak SW lobe was detected more often at 1.7 GHz than at 5 GHz. Its flux density at 5 GHz is about half that at 1.7 GHz.) A typical error is 0.2 mJy. A flux density determination was made every 50 minutes for the VLBA observations and every two hours for the APT and EVN observations. The major flares are indicated: C=core, N=NE lobe, S=SW lobe. The arrows near the abscissa show the times of minimum optical light for the binary system. 7 Fig. 3. -- The Motion of the NE and SW Lobes. (top) 1999 June Observations and (bottom) 1998 February Observations. The plotted points show the separation the NE and SW lobes from the core compo- nent measured every 50 min from the VLBA snaphots at 1.7 and 5.0 GHz and every two hours at 5.0 GHz from the APT and EVN for 1999 June. Many of the indicated one-sigma error bars are smaller than the symbol size. The best linear velocity fits for NE lobe (#1) and lobe (#2) in 1999 June and for the NE lobe in 1998 February are indicated by the solid line. The dashed lines through the SW components are not fits to the data, but 50% of that of the NE lobe at the same time. 8 Fig. 4. -- The Flux Density of the Components in Sco X-1 during the 1999 June Observations After Correction for the Time Delay. The jet speed is βj = 1.0. This plot is similar to Figure 2, but with the NE and SW lobe times modified as discussed in the paper. Several points near the N3, N4 and S4 peaks are given for βj = 0.9. 9
astro-ph/0311409
1
0311
2003-11-17T23:35:06
Year-scale morphological variation of the X-ray Crab Nebula
[ "astro-ph" ]
We present year-scale morphological variations of the Crab Nebula revealed by Chandra X-ray observatory. Observations have been performed about every 1.7 years over the 3 years from its launch. The variations are clearly recognized at two sites: the torus and the southern jet. The torus, which had been steadily expanding until 1.7 years ago, now appears to have shrunk in the latest observation. Additionally, the circular structures seen to the northeast of the torus have decayed into several arcs. On the other hand, the southern jet shows the growth of its overall kinked-structure. We discuss the nature of these variations in terms of the pulsar wind nebula (PWN) mechanism.
astro-ph
astro-ph
Young Neutron Stars and Their Environments IAU Symposium, Vol. 218, 2004 F. Camilo and B. M. Gaensler, eds. Year-scale morphological variation of the X-ray Crab Nebula Koji Mori, David N. Burrows, George G. Pavlov Department of Astronomy and Astrophysics, 525 Davey Laboratory, The Pennsylvania State University, University Park, PA 16802, U.S.A. J. Jeff Hester Department of Physics and Astronomy, Arizona State University, Tempe, AZ 85287-1504, U.S.A. Shinpei Shibata Department of Physics, Yamagata University, Yamagata 990-8560, JAPAN Hiroshi Tsunemi Department of Earth and Space Science, Graduate School of Science, Osaka University, 1-1 Machikaneyama, Toyonaka, Osaka 560-0043 JAPAN Abstract. We present year-scale morphological variations of the Crab Nebula revealed by Chandra X-ray observatory. Observations have been per- formed about every 1.7 years over the 3 years from its launch. The variations are clearly recognized at two sites: the torus and the southern jet. The torus, which had been steadily expanding until 1.7 years ago, now appears to have shrunk in the latest observation. Additionally, the circular structures seen to the northeast of the torus have decayed into several arcs. On the other hand, the southern jet shows the growth of its overall kinked-structure. We discuss the nature of these variations in terms of the pulsar wind nebula (PWN) mechanism. 1. Introduction The 5-month monitoring observations of Chandra and HST showed short-term (days to weeks) morphological variations of the Crab Nebula (Mori et al. 2002; Hester et al. 2002). A variable inner ring and wisps emerging from it visu- ally revealed the existence of the termination shock of the pulsar wind and its downstream flow, respectively. In addition to such short-term, therefore more noticeable, variations, expansion of the outer torus on longer time scale (months) was also discovered. However, this seems to contradict with the fact that the angular extent of the torus is almost constant over 25 years (Mori et al. 2002); if the torus kept expanding at the observed rate, it would have become almost 1 2 Mori. et al. twice larger during this period. On the other hand, the southern jet did not show a strong variation of its overall structure. This differs from the highly variable jet of Vela (Pavlov et al. 2003), which is an another famous example of a variable PWN. Here we present long-term (years) morphological variations of the Crab Nebula which are newly discovered comparing Chandra observations over 3 years. 2. Observations The Crab Nebula has been observed by several PIs with different objectives. The first observation was performed just after launch (Weisskopf et al. 2000) and the latest one was done 3.3 years after that to witness the transit of Titan, Saturn's largest moon, across the Crab Nebula (Mori et al. 2003, in preparation). In this paper, we used those two observations as well as one of the monitoring observations which were performed almost at the midpoint of the first and latest ones. Hereafter, we call them 1st, 2nd, and 3rd epoch observations. They are roughly spaced at 1.7 years interval. Table 1 summarizes these observations. Table 1. Observational logs of Chandra observations used in this paper Notation Date Interval (year)a Exposure (ksec) Configurationb 1st 2nd 3rd 08/29/99 04/06/01 01/05/03 - 1.7 3.3 aAn interval from the 1st observation bNumber in parenthesis indicates the frame time 2.7 2.6 35 ACIS(3.2)+HETG ACIS(0.2) ACIS(0.3)+HETG 3. Result and Discussion In contrast with short-term variations are seen at inner regions, the long-term variations are prominent at outer regions: the torus and the southern jet. In the following sections, we will discuss these variations separately. 3.1. The northeast of the torus Figure 1 shows expanded views of the northeast of the torus. The brightness scales of these images are normalized to the bright region of the torus, not by exposure time, because of the differences of the observational configurations. Images of the 1st and 2nd epoch observations are quite similar. The bright "thick" torus is encircled by circular structures at its northeastern end. As reported previously, these two structures appear to expand between the 1st and 2nd epoch observations (Mori et al. 2000; Mori 2002). In contrast with the similarity between the 1st and 2nd epoch images, the morphological transition from the 2nd to the 3rd epoch image is remarkable. The bright region of the torus appears to have became shrunken and "thin", and the circular structures have decayed into several arcs. Now it is clear that the torus is not steadily expanding; the overall extent is almost constant over decades, but the boundary of the torus varies with an Year-scale variation of the Crab nebula 3 Figure 1. Expanded views around the northeast of the torus of 1st (top left), 2nd (top right), and 3rd (bottom left) epoch observations. Black lines seen in the 3rd epoch image are instrumental effects. It is as if we angular scale of a few arcseconds and a time scale of years. were seeing a top of a fountain. Greiveldinger & Aschenbach (1999) reported surface brightness variations of the Crab Nebula using ROSAT HRI observations spanning 6 years. Although ROSAT could not detect a morphological variation, they showed a monotonic increase of the surface brightness at the northeastern region of the torus. The brightness variation discovered by ROSAT is most likely related to the morphological variation discovered by Chandra, suggesting that the time scale of the variation of the torus might be about a decade rather than a few years. 3.2. The southern jet Figure 2 shows expanded views of the southern jet of the 1st, 2nd, and 3rd epoch observations. Although the displacement of the jet is quite small, the series of images clearly shows the growth of its overall kinked-structure. The variability of the jet is reminiscent of northern variable jet of Vela (Pavlov et al. 2003). Table 2 compares the time scale of the variability and the width between the Crab and Vela Jets. The Crab jet is 10 times larger and varies 10 times slower than the Vela jet. If these variations are due to MHD instability, the time scale is proportional to the Alven crossing time, τ ∼ r/ν, where r is the width of the jet and ν is the Alven velocity (Begelman 1998). Considering that the Alven velocity in an ultrarelativistic plasma like the Crab and Vela PWN is more or less the same, above equation applies to both the Crab and Vela jet. Therefore, 4 Mori. et al. Figure 2. Expanded views of the southern jet of 1st (top left), 2nd (top right), and 3rd (bottom left) epoch observations. The line in each image traces the axis of the jet. All three lines are superimposed in the bottom right image. it is suggested that a common mechanism is responsible for the variability of the Crab and Vela jet. Table 2. Comparison of variable jets of the Crab Nebula and Vela Time scale (day) Width (cm) 2.9 × 1017 3 × 1016 150-500 10-30 Crab Vela References Begelman, M. C. 1998, ApJ, 493, 291 Greiveldinger, C. & Aschenbach, B. 1999, ApJ, 510, 305 Hester, J. J. et al. 2002, ApJ, 577, 871 Kennel, C. F., & Coroniti, F. V. 1984, ApJ, 283, 694 Mori, K. et al. 2002, in Neutron Stars in Supernova Remnants, ed. P. O. Slane & B. M. Gaensler, p157 Mori, K. 2002, Ph. D. Thesis, Osaka University Pavlov, G. G. et al. 2003, ApJ, 591, 1157 Weisskopf, M. et al. 2000, ApJ, 536, 81
astro-ph/0207475
1
0207
2002-07-22T17:58:08
SPI science prospects
[ "astro-ph" ]
With the launch of ESA's INTEGRAL satellite in october 2002, a gamma-ray observatory will become available to the scientific community that combines imaging and spectroscopic capacities in the 20 keV to 10 MeV energy range. In this paper, we summarise the science prospects of the SPI spectrometer aboard INTEGRAL which provides unprecedented spectral resolution with good imaging capabilities and high sensitivity. Emphasise will be given to the key objectives of SPI which are the determination of the galactic positron origin, the study of galactic star formation and nucleosynthesis activity via the radioactive trace isotopes 26Al and 60Fe, and the study of nucleosynthesis in supernova explosions and the gamma-ray line emission from their remnants.
astro-ph
astro-ph
SPI SCIENCE PROSPECTS J. KN ODLSEDER and J.-P. ROQUES Centre d'Etude Spatiale des Rayonnements, B.P. 4346, 31028 Toulouse, FRANCE E-mail: [email protected] With the launch of ESA's INTEGRAL satellite in october 2002, a gamma-ray observatory will become available to the scientific community that combines imaging and spectroscopic capacities in the 20 keV to 10 MeV energy range. In this paper, we summarise the science prospects of the SPI spectrometer aboard INTEGRAL which provides unprecedented spectral resolution with good imaging capabilities and high sensitivity. Emphasise will be given to the key objectives of SPI which are the determination of the galactic positron origin, the study of galactic star formation and nucleosynthesis activity via the radioactive trace isotopes 26Al and 60Fe, and the study of nucleosynthesis in supernova explosions and the gamma-ray line emission from their remnants. 1 The SPI telescope on INTEGRAL Due to continuing progress in instrumentation, the field of gamma-ray line astronomy has become a new complementary window to the universe. With the COMPTEL and OSSE telescopes onboard CGRO, the sky has been imaged for the first time in the light of gamma-ray lines, leading to maps of 511 keV annihilation radiation and 26Al 1.809 MeV emission. New gamma- ray lines have been discovered, such as the 1.157 MeV line from 44Ti or several decay lines from 56Co and 57Co. Gamma-ray lines probe aspects of nucleosynthesis, stellar evolution, and supernova physics that are difficult to access by other means. Additionally, they provide tracers of galactic activity and improve our understanding of the interstellar recycling processes. With the upcoming INTEGRAL satellite, foreseen for launch in october 2002, ESA provides a gamma-ray observatory to the scientific community that combines imaging and spectroscopic capacities in the 20 keV to 10 MeV energy range. INTEGRAL is equipped with two gamma-ray telescopes, optimised for high-resolution imaging (IBIS) and high-resolution spectroscopy (SPI), supplemented by two X-ray monitors (JEM-X) and an optical monitor (OMC). With respect to precedent instruments, the INTEGRAL telescopes provide enhanced sensitivity together with improved angular and spectral resolution. In particular, SPI will map gamma-ray lines with an angular resolution of about 2◦ and a spectral resolution of E/∆E ∼ 500 (at 1 MeV), cor- responding to Doppler velocities of ∼ 600 km s−1. Consequently, gamma-ray line astrophysics figures among the prime objectives of this telescope. SPI consists of a pixelised gamma camera with a geometrical area of ∼ 500 cm−2, made of 19 hexagonal high-purity germanium crystals cooled actively to a temperature of ∼ 85 K. Imaging capabilities are achieved by placing a coded mask at about 1.7 meters above the detector plane, providing the spatial modulation of the incoming gamma-rays that allow the reconstruction of the underlying source intensity distributions. The mask is made of 3 cm thick tungsten elements which provide the necessary absorption efficiency up to high gamma-ray energies. The wide field of view of 16◦ (fully coded; 34◦ partially coded) is defined by an active shield made of BGO scintillator crystals which also acts as anticoincidence shield for instrumental background reduction. A plastic scintillator placed under the mask aims in reducing the 511 keV instrumental background line, improving the telescope's sensitivity at this astrophysically important energy. A pulse shape discrimination electronics provides further background reduction in the 200 keV - 2 MeV domain, which results in an expected (narrow) gamma-ray line sensitivity of 5 × 10−6 ph cm−2s−1 (3σ) at 1 MeV for an observing time of 106 seconds. In the following sections, key science topics that can be addressed with these performances will be presented. Emphasise will be given to gamma-ray line spectroscopy since the SPI tele- scope has been optimised for such studies. However, with its excellent continuum emission sensitivity, in particular in combination with the large field of view and moderate angular reso- lution, SPI is also well suited for studying continuum sources, in particular if they are of diffuse nature. 2 Stellar nucleosynthesis 2.1 Key questions Stars more massive than ∼ 8 M⊙ are the most prolific nucleosynthesis sites in the universe. Dur- ing their short lives they synthesise large quantities of heavy elements that enrich the interstellar medium either through stellar wind ejection or at the final explosion of the star in a supernova event. Although the big picture of element synthesis is already understood since the 50ies 4, many details are still poorly known, and theoretical yield predictions generally suffer from large uncertainties 36. In particular, the physics of mixing processes within massive stars is not well understood, and the impact of stellar rotation and/or close binary evolution can substantially alter abundance patterns 25. By measuring isotopic nucleosynthetic yields using gamma-ray line observations, SPI can provide important constraints on the mixing processes, and may provide clues on the effects of stellar rotation 18. Also, mass loss through stellar winds is crucial for the evolution of the most massive stars, and gamma-ray line observations of mass-losing stars (in particular Wolf-Rayet stars) allow a direct study of its impact on nucleosynthesis yields 29. Similar to the use of radioactive tracers in medicine, freshly produced radioactive isotopes trace also the distribution of nucleosynthesis activity throughout the Galaxy, and gamma-ray line observations may be employed to study this activity. As we will show in the next section, the radioactive isotope 26Al is an excellent candidate tracer for this purpose, complementing other tracers of massive star activity such as the molecular gas distribution, far-infrared emission or free-free emission from the ionised interstellar medium. 2.2 Galactic structure and distribution With the detection and the mapping of the 1.809 MeV gamma-ray line from 26Al, gamma-ray line astronomy has made important progress during the last 15 years. The COMPTEL telescope aboard CGRO has provided the first all-sky map of this radioactive isotope, which with a lifetime of ∼ 106 years is an unambiguous proof that nucleosynthesis is still active in our Galaxy 30. The 1.809 MeV map shows the galactic disk as the most prominent emission feature, demonstrating that 26Al production is clearly a galaxywide phenomenon. The observed intensity profile along the galactic plane reveals asymmetries and localised emission enhancements, characteristic for a massive star population that follows the galactic spiral structure. Thus, 26Al production seems related to the massive star population. Correlation studies using tracer maps for various source candidate populations strongly support this suggestion, which in particular revealed that 1.809 MeV gamma-ray line emission follows closely the distribution of galactic free-free emission 20. This suggests which is powered by the ionising radiation of stars with initial masses > 20 M⊙ 0.012 0.010 0.008 0.006 0.004 0.002 ) 2 - c p k O • M ( y t i s n e d e c a f r u s s s a M 0.000 0 10 2 Galactocentric distance (kpc) 4 6 8 12 Figure 1: Radial 26Al density profile (left) and COMPTEL 1.809 MeV image of the Vela region (right). The filled circles compare the angular resolution of the COMPTEL and the SPI telescopes. that explosive 26Al production in supernovae may be less important than previously thought and hydrostatic nucleosynthesis in massive mass-losing stars may possibly be the primary production channel for galactic 26Al. Having established the correlation between 1.809 MeV emission and massive star populations, 26Al becomes an excellent tracer of recent galactic star formation. By refining the knowledge about the 1.809 MeV emission distribution, SPI will provide a unique view on the star formation activity in our Galaxy. To illustrate this potential, the radial 26Al mass density distribution as derived by COMPTEL is shown in the left panel of Fig. 1. The bulk of galactic star formation occurs at distances of less than 6 kpc from the galactic centre. Star formation is also present within the central 3 kpc of the Galaxy, although at a poorly determined rate. There are in- dications for enhanced star formation between 3 − 6 kpc, coinciding with the molecular ring structure. Enhanced star formation is also seen in the solar neighbourhood (8 − 9 kpc) which probably corresponds to activity in the local spiral arm. The radial 26Al profile is probably not directly proportional to the radial star formation pro- file since 26Al nucleosynthesis may depend on metallicity 26. It will be important to determine this metallicity dependence in order to extract the true star formation profile from gamma-ray line data. Valuable information about the metallicity dependence will come from a precise deter- mination of the 1.809 MeV longitude profile by SPI, and the comparison of this profile to other tracers of star formation activity, such as galactic free-free emission. Additionally, observations of gamma-ray lines from 60Fe, an isotope that is mainly believed to be produced during super- nova explosions, may help to distinguish between hydrostatically and explosively produced 26Al, and therefore may allow disentangling the metallicity dependencies for the different candidate sources. A precise determination of the 1.809 MeV latitude profile by SPI will provide important information about the dynamics and the mixing of 26Al ejecta within the interstellar medium. High velocity 26Al has been suggested by measurements of a broadened 1.809 MeV line by the GRIS spectrometer 28, although this observation is at some point at odds with the earlier observation of a narrow line by HEAO 3 24. In any case, the propagation of 26Al away from its origin should lead to a latitude broadening with respect to the scale height of the source population, and the observation of this broadening may allow the study of galactic outflows CepOB2 NGC7160 NGC7261 Tr37 NGC7235 NGC7128 NGC7380 Ber96 CepOB1 Ber94 NGC7067 IC5146 LacOB1 ) g e d ( e d u t i t a L 10 5 0 -5 -10 7 NGC6910 Ber86 IC4996 8 3 1 NGC6871 VulOB1 9 2 NGC6913 Biu2 Ros5 Ber87 Ros4 NGC6883 Rup175 NGC6823 ) g e d ( e d u t i t a L 10 10 5 5 0 0 -5 -5 -10 -10 100 90 80 Longitude (deg) 70 100 100 80 90 90 80 Longitude (deg) 70 70 Figure 2: Young open clusters (asterisks) and OB associations (circles) in the Cygnus region (left) and 1.809 MeV COMPTEL image (contours) superimposed on a 53 GHz free-free emission map of the region (right) (from Knodlseder et al. 2002). and the mass transfer between disk and halo of the Galaxy. Actually, COMPTEL 1.809 MeV observations restrict the scale height of the galactic 26Al distribution to z < 220 pc 7, which certainly excludes a ballistic motion of 26Al at a speed of 500 km s−1. The excellent energy resolution of SPI will easily allow to decide whether the 1.809 MeV line is broadened or not, and the improved angular resolution and sensitivity with respect to COMPTEL will allow to determine the scale height of the galactic 26Al distribution much more precisely. The expected energy resolution of SPI of ∼ 2.5 keV at 1.809 MeV converts into a velocity resolution of ∼ 400 km s−1, allowing for line centroid determinations of the order of 50 km s−1 for bright emission features. Thus, in the case of an intrinsically narrow 1.809 MeV line, line shifts due to galactic rotation should be measurable by SPI 9. Although this objective figures certainly about the most ambitious goals of SPI observations, a coarse distance determination of 1.809 MeV emission features based on the galactic rotation curve seems in principle possible. 2.3 Massive star clusters Complementary to the study of the large-scale distribution of the 1.809 MeV emission by SPI will be observations of nearby, localised 1.809 MeV emission regions, such as Vela, Cygnus, Carina, or Orion. The aim of these observations will be the identification of emission counterparts at other wavelengths in order to associate the nucleosynthesis activity to individual objects or specific groups such as OB associations or young open clusters. Already with COMPTEL, such studies have proven to provide important insights into the nature of 26Al sources, and in constraining nucleosynthetic yields for individual objects 18. To illustrate the potential of SPI, a contour map of the 1.809 MeV emission in the Vela region obtained by COMPTEL is shown in the right panel of Fig. 1. There is a wealth of potential 26Al sources in this field, but the limited angular resolution of COMPTEL does not allow for a clear identification of the dominant contributors. Additionally, the sensitivity of COMPTEL is not sufficient to clearly separate diffuse from point-like emission, leading to an additional uncertainty in the association of emission structures with 26Al sources. With improved sensitivity and angular resolution, SPI will help to overcome this problem. Deep exposures of localised emission features will sufficiently constrain the 1.809 MeV morphology to associate the structure with candidate sources in the field. In the Vela region, which is part of the INTEGRAL core program, a detection of the Wolf-Rayet star WR 11 is awaited, and the contributions of the Vela SNR, the RX J0852.0-4622 supernova remnant (Vela Junior), and OB star associations (associated to the radio source RCW 38) should be measurable. In the Cygnus region, 26Al 1.809 MeV gamma-ray line emission from the massive young globular cluster Cyg OB2 should be detectable by SPI, allowing for the first time the study of nucleosynthesis in an individual massive star association 18. Indeed, the Cygnus region houses a variety of young stellar associations and clusters within an area of ∼ 10◦ in diameter, separated by a few degrees (see Fig. 2). The SPI angular resolution is well matched to disentangle the gamma-ray line emission from these individual stellar groups, and may even allow to trace the propagation of the nucleosynthesis ejecta into the surrounding interstellar medium. The actual COMPTEL observations (see Fig. 2) indeed show emission maxima close to the Cyg OB2 cluster, yet the limited photon statistics do not allow for a detailed morphological study of this emission. With its enhanced sensitivity and angular resolution, SPI should provide a much more detailed image of this region. In particular, the correlation of 1.809 MeV emission and free-free emission from the ionised interstellar medium may be studied in greater detail, providing a more comprehensive picture of the interplay between massive stars and the surrounding interstellar medium. 3 Supernova nucleosynthesis 3.1 Supernovae Supernovae are the most prolific nucleosynthesis sites in the Universe, producing a large variety of chemical elements that are ejected into the interstellar medium by the explosion. Among those are radioactive isotopes that decay under gamma-ray line emission with lifetimes that are sufficiently long to allow escape in regions that are transparent to gamma-rays. In particular, substantial amounts of 56Ni and 57Ni are produced which subsequently decay under gamma-ray line emission to 56 57Co and finally to 56 , , 57Fe. Observationally, type Ia events are easier to observe than the other supernova classes because they produce an order of magnitude more radioactive 56Ni than the other types, and because they expand rapidly enough to allow the gamma-rays to escape before all the fresh radioactivity has decayed. From the SPI sensitivity and the observed type Ia supernova rates together with standard models of type Ia nucleosynthesis, one may estimate the maximum detectable distance for a type Ia event to about 15 Mpc and the event frequency to one event each 5 years 35. Hence, during an extended mission lifetime of 5 years, SPI has statistically spoken the chance to detect one such event. Nevertheless, CGRO observations have taught us that supernovae are intriguing objects, and even at the detection threshold, their observation may provide interesting implications on the progenitor nature or explosion mechanism. For example, Morris et al 27 report the detection of the unusually bright SN 1991T in NGC 4527 (at a distance between 13−17 Mpc) by COMPTEL, implying a 56Ni mass of 1.3 − 2.3 M⊙. This mass exceeds all theoretical expectations, requiring possibly a super-Chandrasekhar scenario to explain the observations. Indeed, high (∼ 1 M⊙) 56Ni masses have also been inferred from optical spectra 33 32, suggesting a white-dwarf merger at the origin of the explosion 8. , Another interesting example is SN 1998bu in M96, which shows the characteristics of a rather typical type Ia event at a distance of about 11 Mpc. Theoretical nucleosynthesis models predict that the radioactive decay of 56Co in SN 1998bu should lead to a peak flux of (1 − 5) × 10−5 ph cm−2s−1 in the 847 keV line 11. However, SN 1998bu was observed by OSSE for over 140 days and by COMPTEL for almost 90 days without any positive detection 21 10. The upper time- averaged 847 keV flux limit of OSSE amounts to 3 × 10−5 ph cm−2s−1, the COMPTEL limit of 4 × 10−5 ph cm−2s−1 is comparable. For the 1.238 MeV line, COMPTEL imposes an even more stringent flux limit of 2 × 10−5 ph cm−2s−1. Thus, the observations start to constrain type Ia supernova models, excluding for example the Helium cap model for SN 1998bu 10. Observing SN 1998bu by SPI would probably have been a major breakthrough for observational gamma-ray , line astrophysics. Even if the 847 keV line would have been broadened to 50 keV, which is probably rather pessimistic, SPI would have achieved a sensitivity of 10−5 ph cm−2s−1 within a comparable exposure time (∼ 100 days). At this level, either SN 1998bu would have been clearly detected or the non-detection would have ruled out all existing thermonuclear supernova models. Assuming that SN 1998bu was indeed close to detection (at a 847 keV flux of say 3 × 10−5 ph cm−2s−1), SPI would have detected the 847 keV line at a significance of about 10σ, allowing for valuable line profile studies. 3.2 SN 1987A The explosion of SN 1987A in the Large Magellanic Cloud was a great opportunity for gamma- ray line astronomy. For the first time, a supernova explosion occurred close enough to be in reach of available gamma-ray telescopes. The direct observation of the gamma-ray lines from 56Co and 57Co in SN 1987A was a brilliant confirmation of supernova theory which explains lightcurve characteristics as result of the radioactive decay of these isotopes. The observed relative intensities of the gamma-ray lines from 56Co and 57Co indicated a 57Ni/56Ni ratio between 1.5 − 2 times the solar ratio of 57Fe/56Fe, consistent with core collapse supernova models 37. Surprisingly, the 56Co lines were detected already 6 months after explosion, at an epoch where standard onion-shell supernova expansion models still predicted a substantial gamma-ray opacity for the envelope. The gamma-ray line lightcurves presented clear evidence that 56Co was found over a large range of optical depths, with a small fraction at very low depth 22. Probably some fragmentation of the ejecta and acceleration of the emitting radioactivity are required to explain the observations. The acceleration hypothesis is supported by various gamma-ray line profile measurements all indicating line widths of order 1 % FWHM, corresponding to Doppler velocities of 3000 km s−1. Measurements of the bolometric SN 1987A lightcurve indicate that also ∼ 10−4 M⊙ of 44Ti have been produced during the explosion, and the observation of the corresponding radioactive decay lines at 67.9, 78.4, and 1157 keV by SPI presents a unique tool to probe core collapse physics. 44Ti is synthesised in the innermost layers of the star during the so-called α-rich freeze- out which is sensitive to entropy, and the determination of the 44Ti yield would provide a direct measure of this quantity. Also, the gamma-ray line profile carries valuable information about the velocity distribution of the matter close to the compact remnant, hence 44Ti observations inform us about the explosion dynamics and possible ejecta acceleration. Radioactive decay of 60Co, accompanied by gamma-ray line emission at 1.173 and 1.332 MeV, is sensitive to the neutron excess in the supernova, providing a unique chance for a direct measurement of this important parameter in SN 1987A. The measurement of the γ-ray line fluxes will allow the determination of the 44Ti and 60Co yields in SN 1987A, providing information about the position of the mass- cut, the maximum temperature and density reached during the passage of the shock wave in the ejecta, and the neutron excess. 3.3 Supernova remnants Until recently, the census of recent galactic supernova events was exclusively based on historic records of optical observations and amounted to 6 events during the last 1000 years. Due to galactic absorption and observational bias, this census is by far not complete. At gamma-ray energies, however, the Galaxy is transparent, and hence gamma-ray line observations of the 44Ti isotope have the potential to unveil yet unknown young supernova remnants. The proof of principle was achieved by the observation of a 1.157 MeV gamma-ray line from the 320 years old Cas A supernova remnant using the COMPTEL telescope 16. Evidence for another galactic 44Ti source has been found in the Vela region where no young supernova remnant was known before 15. Triggered by this discovery, unpublished ROSAT X-ray data showing a spherical structure at the position of the new 44Ti source were reconsidered and lead to the discovery of a new supernova remnant, now called RX J0852.0-4622 2 (also named Vela Junior). In the meanwhile, the remnant has also been discovered at radio wavelengths 6. Although the 44Ti observation is only marginal, it is the first time that gamma-ray line observations triggered the discovery of a new supernova remnant. From the X-ray data, an age of less than 1500 yrs and a distance < 1 kpc has been inferred. Adding the 44Ti observations further constrains the age and distance to ∼ 680 yrs and ∼ 200 pc, respectively 1. Interestingly, nitrate abundance data from an Antarctic ice core provide evidence for a nearby galactic supernova 680 ± 20 years ago, compatible with the 44Ti data 5. Given the marginal detection of the 1.157 MeV line from RX J0852.0-4622, a confirmation by SPI will be crucial for the further understanding of this object. 44Ti line-profile measurements will provide complementary information on the expansion velocity and dynamics of the innermost layers of the supernova ejecta. The regular galactic plane scans and the deep exposure of the central radian that INTEGRAL will perform during the core program of the mission will lead to a substantial build-up of exposure time along the galactic plane, enabling the detection of further hidden young galactic supernova remnants through 44Ti decay. The observed supernova statistics may then set interesting constraints on the galactic supernova rate and the 44Ti progenitors. Indeed, actual observations already indicate that some of the galactic 44Ca may have been produced by a rare type of supernova (e.g. Helium white dwarf detonations) which produces very large amounts of 44Ti 34. 4 Positron annihilation The 511 keV gamma-ray line due to annihilation of positrons and electrons in the interstel- lar medium has been observed by numerous instruments 14. At least two galactic emission components have been identified so far: an extended bulge component and a disk component. Indications of a third component situated above the galactic centre have been reported 31 14, yet still needs confirmation by more sensitive instruments. , The galactic disk component may be explained by radioactive positron emitters, such as 26Al, 44Sc, 56Co, and 22Na. The origin of the galactic bulge component is much less clear. An apparent 511 keV flux variation from the galactic centre has led to the idea that a compact object might be responsible for the galactic bulge emission, yet the flux variation has turned out to be insignificant 23. Also, contemporaneous observations with the SMM satellite and latest observations by OSSE and TGRS show no evidence for time-variability. This limits the flux level of any variable 511 keV point source to less than 4 × 10−4 ph cm−2s−1. On the other hand, the observation of broadened and red-shifted annihilation features from 1E 1740.7-2942 3 and Nova Muscae 12 has been considered as evidence for positron production in compact objects. However, contemporaneous observation by OSSE and SIGMA of an outburst of 1E 1740.7-2942 in September 1992 gave contradictory results 17, casting some doubt on the contribution of compact objects to the galactic positron budget. The origin of the galactic bulge positrons will be one of the key-questions addressed by SPI. Narrow-line transient features with fluxes of 4 × 10−4 ph cm−2s−1 should be detectable by SPI within less than one hour. If the feature is broadened by 300 keV (as observed for example in 1E 1740.7-2942) the required observation time increases to ∼ 12 hours. Hence, the weekly galactic plane scan together with the central radian deep exposure performed during the INTEGRAL core program will provide a unique survey of the galactic bulge, capable of detecting even faint transient 511 keV emission events. SPI will also provide a detailed map of 511 keV emission from the Galaxy. Using this map, the morphology of the galactic bulge can be studied in detail, and the question on the contribution of point sources to the galactic bulge emission can be addressed. In particular, the ratio between bulge and disk emission, which is only poorly constrained by existing data, will be measured more precisely, allowing for more stringent conclusions about the positron sources of both components. The 511 keV map will also answer the question about the reality of the positive latitude enhancement, which may provide interesting clues on the activity near to the galactic nucleus (see von Ballmoos, these proceedings). The 511 keV line shape carries valuable information about the annihilation environment which will be explored by SPI. The dominant annihilation mechanism sensitively depends on the temperature, the density, and the ionisation fraction of the medium, and the measurement of the 511 keV line width allows the determination of the annihilation conditions 13. Observations of a moderately broadened 511 keV line towards the galactic centre indicate that annihilation in the bulge mainly occurs in the warm neutral or ionised interstellar medium 14. By making spatially resolved line shape measurements, SPI will allow to extend such studies to the entire galactic plane, complementing our view of galactic annihilation processes. With its good continuum sensitivity, SPI will also be able to detect the galactic positronium continuum emission below 511 keV. The intensity of this component with respect to that of the 511 keV line carries complementary information about the fraction f of annihilations via positronium formation, probing the thermodynamic and ionisation state of the annihilation environment 13. 5 Conclusions With the research topics that we detailed in the above sections, the list of SPI science prospects is by far not complete 19. Further gamma-ray line sources such as novae, accreting black holes, or nuclear interactions of cosmic-rays with the interstellar medium are valuable targets of SPI which promise new insights into nucleosynthesis processes and particle acceleration. Since SPI is the spectrometer on INTEGRAL we deliberately excluded topics related to continuum emission -- those will be addressed by the corresponding paper of the IBIS collaboration (Lebrun, these proceedings). Nevertheless, topics like the diffuse gamma-ray background, the diffuse galactic gamma-ray emission, pulsars, active galactic nuclei, or gamma-ray bursts, are equally important for SPI, which has a continuum sensitivity comparable to IBIS, but which is optimised for large-scale emission (IBIS provides a much better angular resolution but is less sensitive to extended emission features). On the other hand, IBIS may also detect gamma-ray lines and can help to identify counterparts by means of the high localisation accuracy. Thus, SPI and IBIS are complementary instruments onboard INTEGRAL, which, when combined, will explore the gamma-ray sky far beyond the established horizon. References 1. B. Aschenbach, A.F. Iyudin and V. Schonfelder, A&A 350, 997 (1999). 2. B. Aschenbach, Nature 396, 141 (1998). 3. L. Bouchet, P. Mandrou, J.-P. Roques, et al., ApJ 383, L45 (1991). 4. E.M. Burbidge, G.R. Burbidge, W.A. Fowler and F. Hoyle, Reviews of Modern Physics 29, 547 (1957). 5. C.P. Burgess and K. Zuber, New Scientist 163, 2204 (1999). 6. J.A. Combi, G.E. Romero and P. Benaglia, ApJ 519, L177 (1999). 7. R. Diehl, U. Oberlack, J. Knodlseder, et al., in: Proc. 4th Compton Symposium, eds. C.D. Dermer, M.S. Strickman and J.D. Kurfess, 1114 (1997). 8. A. Fisher, D. Branch, K. Hatano and E. Baron, MNRAS 304, 67 (1999). 9. N. Gehrels and W. Chen, A&AS 120C, 331 (1996). 10. R. Georgii, S. Pluschke, R. Diehl, et al., in: Proc. 5th Compton Symposium, eds. M.L. McConnell and J.M. Ryan, 49 (2000). 11. J. G´omez-Gomar, J. Isern and P. Jean, MNRAS 295, 1 (1998). 12. A. Goldwurm, J. Ballet, B. Cordier, et al., ApJ 389, L79 (1992). 13. N. Guessoum, R. Ramaty, R.E. Lingenfelter, ApJ 378, 170 (1991). 14. M.J. Harris, B.J. Teegarden, T.L. Cline, et al., ApJ 501, L55 (1998). 15. A.F. Iyudin, V. Schonfelder, K. Bennett, et al., Natur 396, 142 (1998). 16. A.F. Iyudin, R. Diehl, H. Bloemen, et al., A&A 284, 1 (1994). 17. G.V. Jung, D.J. Kurfess, W.N. Johnson, et al., A&A 295, L23 (1995). 18. J. Knodlseder, M. Cervino, J.-M. LeDuigou, et al., A&A, in press. 19. J. Knodlseder and G. Vedrenne, in: Proc. of the 4th INTEGRAL Workshop, eds. A. Gimenez, V. Reglero, & C. Winkler ESA SP-459, 23 (2000). 20. J. Knodlseder, K. Bennett, H. Bloemen, et al., A&A 344, 68 (1999). 21. M.D. Leising, L.-S. The, P. Hoflich, et al., AAS HEAD meeting 31, 08.03 (1999). 22. M.D. Leising and G.H. Share, ApJ 357, 638 (1990). 23. W.A. Mahoney, J.C. Ling and W.A. Wheaton, ApJS 92, 387 (1994). 24. W.A. Mahoney, J.C. Ling, W.A. Wheaton, et al., ApJ 286, 578 (1984). 25. G. Meynet and A. Maeder, A&A 361, 101 (2000). 26. G. Meynet, ApJS 92, 441 (1994). 27. D.J. Morris, K. Bennett, H. Bloemen, et al., in: 17th Texas Symposium, N.Y. Acad. Sci. 759, 397 (1995). 28. J.E. Naya, S.D. Barthelmy, L.M. Bartlett, et al., Natur 384, 44 (1996). 29. U. Oberlack, U. Wessolowski, R. Diehl, et al., A&A 353, 715 (2000). 30. U. Oberlack, K. Bennett, H. Bloemen, et al., A&A 120C, 311 (1996). 31. W.R. Purcell, L.-X. Cheng, D.D. Dixon, et al., ApJ 491, 725 (1997). 32. P. Ruiz-Lapuente, E. Cappellaro, M. Turatto, et al., ApJ 387, L33 (1992). 33. J. Spyromilio, W.P.S. Meikle, D.A. Allen and J.R. Graham, MNRAS 258, 53 (1992). 34. L.-S. The, R. Diehl, D.H. Hartmann, et al., in: Proc. 5th Compton Symposium, eds. Mark L. McConnell and James M. Ryan, 64 (1999). 35. F.X. Timmes and S.E. Woosley, ApJ 489, 160 (1997). 36. S.E. Woosley, in: Astronomy with Radioactivities, eds. R. Diehl and D. Hartmann, 133 (1999). 37. S.E. Woosley and R.D. Hoffman, ApJ 368, 31 (1991).
0704.3088
1
0704
2007-04-23T20:59:04
The 74MHz System on the Very Large Array
[ "astro-ph" ]
The Naval Research Laboratory and the National Radio Astronomy Observatory completed implementation of a low frequency capability on the VLA at 73.8 MHz in 1998. This frequency band offers unprecedented sensitivity (~25 mJy/beam) and resolution (~25 arcsec) for low-frequency observations. We review the hardware, the calibration and imaging strategies, comparing them to those at higher frequencies, including aspects of interference excision and wide-field imaging. Ionospheric phase fluctuations pose the major difficulty in calibrating the array. Over restricted fields of view or at times of extremely quiescent ionospheric ``weather'', an angle-invariant calibration strategy can be used. In this approach a single phase correction is devised for each antenna, typically via self-calibration. Over larger fields of view or at times of more normal ionospheric ``weather'' when the ionospheric isoplanatic patch size is smaller than the field of view, we adopt a field-based strategy in which the phase correction depends upon location within the field of view. This second calibration strategy was implemented by modeling the ionosphere above the array using Zernike polynomials. Images of 3C sources of moderate strength are provided as examples of routine, angle-invariant calibration and imaging. Flux density measurements indicate that the 74 MHz flux scale at the VLA is stable to a few percent, and tied to the Baars et al. value of Cygnus A at the 5 percent level. We also present an example of a wide-field image, devoid of bright objects and containing hundreds of weaker sources, constructed from the field-based calibration. We close with a summary of lessons the 74 MHz system offers as a model for new and developing low-frequency telescopes. (Abridged)
astro-ph
astro-ph
The 74 MHz System on the Very Large Array N. E. Kassim, T. Joseph W. Lazio Naval Research Laboratory, Remote Sensing Division, Code 7213, Washington, DC 20375-5351 [email protected] [email protected] W. C. Erickson U. of Tasmania, School of Math. & Physics, G.P.O. Box 252-21, Hobart, Tasmania 7001, Australia National Radio Astronomy Observatory, P.O. Box O, Socorro, NM 87801 R. A. Perley [email protected] W. D. Cotton National Radio Astronomy Observatory, 520 Edgemont Road, Charlottesvil le, VA 22903 [email protected] E. W. Greisen National Radio Astronomy Observatory, P.O. Box O, Socorro, NM 87801 [email protected] A. S. Cohen, B. Hicks Naval Research Laboratory, Remote Sensing Division, Code 7213, Washington, DC 20375-5351 [email protected] [email protected] H. R. Schmitt – 2 – Naval Research Laboratory, Remote Sensing Division, Code 7215, Washington, DC 20375-5351 and Interferometrics, Inc., 13454 Sunrise Val ley Drive, Suite 240, Herndon, VA 20171 [email protected] and D. Katz U.S. Naval Academy, Physics 9C, Annapolis, MD21402-5026 [email protected] ABSTRACT The Naval Research Laboratory and the National Radio Astronomy Observa- tory completed implementation of a low frequency capability on the Very Large Array at 73.8 MHz in 1998. This frequency band offers unprecedented sensitivity (∼ 25 mJy beam−1 ) and resolution for low-frequency observations. The longest baselines in the VLA itself provide 25′′ resolution; the system has recently been extended to the nearby Pie Town antenna of the Very Long Baseline Array, which provides resolutions as high as 12′′ . This paper reviews the hardware, the calibration and imaging strategies of this relatively new system. Ionospheric phase fluctuations pose the ma jor difficulty in calibrating the array, and they influence the choice of calibration strategy. Over restricted fields of view (e.g., when imaging a strong source) or at times of extremely quiescent ionospheric “weather” (when the ionospheric isoplanatic patch size is larger than the field of view), an angle-invariant calibration strategy can be used. In this approach a single phase correction is devised for each antenna, typically via self-calibration; this approach is similar to that used at higher frequencies. Over larger fields of view or at times of more normal ionospheric “weather” when the ionospheric isoplanatic patch size is smaller than the field of view, we adopt a field-based strategy in which the phase correction depends upon location within the field of view. In practice we have implemented this second calibration strategy by modeling the ionosphere above the array using Zernike polynomials. Images of 3C sources of moderate strength are provided as examples of routine, angle- invariant calibration and imaging. Flux density measurements of a sub-sample – 3 – of these sources with previously well determined low frequency spectra indicate that the 74 MHz flux scale at the Very Large Array is stable to a few percent, and that flux densities tied to the Baars et al. value of Cygnus A are reliable to at least 5 percent. We also present an example of a wide-field image, devoid of bright ob jects and containing hundreds of weaker sources, constructed from the field-based calibration. The paper also reviews other practical aspects of low frequency observations, in so far as they differ from those encountered at higher frequencies, including aspects of interference excision and wide-field imaging. We close with a summary of lessons the 74 MHz system offers as a model for new and developing low-frequency telescopes. Subject headings: instrumentation: interferometers — techniques: high angular resolution — techniques: image processing — techniques: interferometric — astrometry 1. Introduction Radio astronomy began with the discovery of celestial radio emission by K. Jansky at 20.5 MHz (Jansky 1935). Throughout the 1950s and 1960s, key discoveries and technolog- ical advances in radio astronomy—at low-frequencies (ν . 100 MHz) in particular—helped form the basis of modern astronomy, including: • The introduction of non-thermal processes as an astrophysical source of emission (Alfv`en & Herlofson 1950; Kiepenheuer 1950; Shkolvsky 1952), motivated by early ob- servations of the diffuse Galactic radio emission at 160 MHz (Reber 1940), though it took some time for the importance of non-thermal processes to be accepted widely; • The discovery of pulsars through observations at 81 MHz (Hewish et al. 1968); and • The development of aperture synthesis interferometry at 38 MHz (Ryle & Vonberg 1946; Ryle et al. 1950; Mills 1952; Ryle 1952). (See also the work by Pawsey et al. (1946) and McCready et al. (1947) using a “sea interferometer” at frequencies be- tween 75 and 3000 MHz.) By the late 1960s and early 1970s, however, interest turned to obtaining high-resolution images. Powerful centimeter-wavelength interferometers began to provide sub-arcminute angular resolution (or sub-arcsecond resolution in the case of the VLA by the late 1970s) with dynamic ranges of several hundred or better. – 4 – Even so, a number of low-frequency interferometers continued to be constructed, some with truly impressive collecting areas (& 105 m2 ), including the UTR-2 (Braude et al. 1978), the Culgoora Radioheliograph, the Cambridge Synthesis Telescope, and the Clark Lake TPT Radio Telescope (Erickson et al. 1982). However, with the exception of the MERLIN (Thomasson 1986; Leahy, Muxlow & Stephens 1989) and GMRT 151 MHz systems (Swarup 1990) and for all telescopes operating below 100 MHz, the maximum baselines were relatively limited (. 5 km). The corresponding angular resolution was relatively poor (& 10′ ), and the resulting high confusion levels meant poor sensitivities (& 1 Jy). The primary constraint on baseline length were the phase distortions imposed by the Earth’s ionosphere over the intrinsically wide fields of view and the lack of suitable algorithms to compensate for the distortions. On baselines longer than a few kilometers, ionospheric phase distortions are se- vere enough to cause decorrelation, making higher-resolution imaging difficult to impossible, especially at the lowest frequencies (≤ 100 MHz). This paucity of large aperture, high-sensitivity, synthesis instruments operating be- low 100 MHz has left this portion of the radio spectrum poorly explored. Yet, there are a variety of topics that could be addressed by a sensitive, high-angular-resolution, low- frequency telescope including: • Continuum spectra over much larger frequency dynamic ranges for studies of shock acceleration and spectral aging in Galactic (supernova remnants) and extragalactic (radio galaxies and galaxy cluster relics) sources; • Efficient detection of large numbers of steep spectrum sources, which can be imaged in some cases, including high-redshift radio galaxies, shocks driven by infalling matter in clusters of galaxies, and pulsars in the Milky Way and possibly in external galaxies; • Probing the ionized interstellar medium (ISM) via measurements of radio-wave scat- tering and absorption, the distribution of low-density ionized gas toward nonthermal sources, and hydrogen and carbon recombination line observations of very high Ryd- berg state atoms; • The large opacity of H II regions below 100 MHz can enable distance determinations to various foreground ob jects, in both the Galaxy and external galaxies, from which the three-dimensional distribution and spectrum of cosmic-ray emissivity can be de- termined as well as being used to measure their emission measures, temperatures, pressures, and ionization states • Detection of coherent emission from sources such as the Sun, Jupiter, pulsars, and possibly radio bursts from nearby stars and extrasolar planets. – 5 – While earlier experience with MERLIN and the VLA (and more recently with the GMRT) demonstrated that large (> 5 km) interferometers could compensate for ionospheric effects below ∼330 MHz, a prototype 74 MHz system was the first to demonstrate that self-calibration techniques can correct for the large ionospheric-induced phase errors below 100 MHz. The prototype 74 MHz system consisted of eight of the VLA’s 28 antennas equipped with 74 MHz receivers, and Kassim et al. (1993) were able to produce images with sub-arcminute resolutions and sub-Jansky sensitivities, thereby demonstrating that these self-calibration techniques were able to correct for ionospheric phase errors on baselines at least as long as the longest VLA baselines (35 km) and, in principle, much longer. Because of the limited number of antennas, the prototype 74 MHz system had relatively poor u-v coverage, so only the strongest (≥ 500 Jy) sources, numbering a dozen or so, were imaged. In 1998 January all 28 VLA antennas were equipped with 74 MHz antennas. This improved the capability of the instrument greatly, and it is now possible to detect hundreds of sources in single fields at high angular resolution and sensitivity. At the same time, a number of innovative procedures and software solutions have been developed to handle the data, some of which are significantly different than at centimeter wavelengths and others of which are applicable to, and new to, centimeter wavelengths. Figures 1 and 2 illustrate the levels of resolution and sensitivity now possible with this new system, improvements that are all the more impressive given that the relatively modest collecting area and low efficiency of the VLA (≃ 2 × 103 m2 effective collecting area, . 10% of many of the other telescopes shown). This paper describes the fully operational 74 MHz system (hardware and software) on the VLA. In §2 we summarize briefly the characteristics of the prototype array as they relate to the current system. In §3 we review general characteristics of the current low frequency system, summarize its general performance, and highlight those aspects that are significantly different than at centimeter wavelengths. We describe the calibration of 74 MHz observations in §4 and their imaging in §5. In §6 we suggest how the current system could be improved via dynamic scheduling. In §7 we discuss possible future expansions of low-frequency synthesis instruments using the lessons from the 74 MHz system, and we present our conclusions in §8. In Appendix A we present selected examples of imaging of moderately strong 3C sources. Throughout the paper we shall illustrate various effects with images or other figures pro- duced from 74 MHz observations. The examples we show are a heterogeneous lot, resulting from a number of different observations of different sources acquired for different purposes. Our ob jective is to present a representative sample of various effects, but not all effects will necessarily be present in every observation. – 6 – 2. Low-Frequency Systems on the VLA The original design of the Very Large Array included only four frequency bands, centered near wavelengths of 21, 6, 2, and 1.3 cm (Napier et al. 1983). However, there is no fundamen- tal reason a low-frequency system cannot operate on the array—the principles of aperture synthesis are as applicable to 50 MHz as they are to 5 GHz. More importantly, the key components of the array—the signal collection (antennas), signal transmission (waveguide), and signal processing (correlator and post-processing)—are essentially frequency indepen- dent within the radio part of the spectrum. As soon as the construction phase of the VLA ended, discussions on implementing a low-frequency capability began. Perley & Erickson (1984) advocated a free-standing array that would make use of the VLA’s infrastructure (most importantly, the waveguide transmission system) to achieve ap- proximately 25′′ resolution. However, no source of funding was obvious, and it was decided subsequently that trial systems could be implemented on the VLA itself to address key questions regarding the calibration and imaging of low frequency, long-baseline data. The initial low-frequency system, operating at 90 cm (300 to 340 MHz), was installed between 1983 and 1989. It is a prime-focus system, as it is impractical to implement a secondary focus system at such a low frequency. The feed is a crossed dipole, situated in front of the Cassegrain subreflector, which thus acts as a (rather imperfect) ground plane. Because of this, and because the phase center is located approximately 50 cm (≈ λ/2) in front of the true focus, this system has both a low efficiency (less than 40%) and a very broad shoulder of width approximately 12◦ in the antenna power pattern. Nevertheless, it has been a very successful and widely used frequency band at the VLA. Most importantly, it encouraged the development of the multi-faceted imaging algorithms (Cornwell & Perley 1992) needed for wide-field, low-frequency observing, as described later in this paper. Its operation also demonstrated the robustness of angle-independent self-calibration (§4.4) for removing ionospheric distortions across the large (∼2.5 ◦FWHP) 90 cm field of view. The success of the 330 MHz system soon led to consideration of a lower frequency facility. A protected radio astronomy frequency allocation exists between 73 and 74.6 MHz. Again, funding constraints led to the decision to deploy a trial system, comprising a simple feed system on a few of the VLA’s antennas. The feed system chosen is essentially the same as that used at 330 MHz—crossed dipoles in front of the subreflector. Because of the long wavelength, the defocussing errors that affect 330 MHz performance severely are not serious at 74 MHz. However, because the antenna itself is only approximately 6λ in diameter, the subreflector is an imperfect ground plane, and the profound effect of the antenna quadrupod structure, it was anticipated that the – 7 – forward gain and sidelobe structure would be fairly poor—as subsequent measurements have borne out. Kassim et al. (1993) describe the prototype 74 MHz system in more detail and describe the initial data calibration and imaging methodologies. 3. The 74 MHz System on the VLA Amplifiers and feeds for the complete 74 MHz system were built during the summer of 1997 by two of us (WCE and BH) at the NRL and deployed during the fall of that year. All antennas were equipped with dipoles by 1998 January. Because of concerns about blockage at higher frequencies, a deployable crossed-dipole feed was designed. The half-wavelength dipoles contribute to blockage and a higher system temperature, resulting in a total sensitivity loss of about 6% at 1.4 GHz and smaller losses at higher frequencies, so they are deployed only during a fraction of the time in each config- uration. A simple mounting system is used—two ropes, each of which supports one dipole, are threaded through eyebolts located on opposite quadrupod legs at the appropriate height. The ends of these ropes are tied to cleats located at a convenient height on the quadrupod legs. The signal cables drop about 7 m to the antenna surface, where they pass through the roof of the vertex room to the amplifiers. Figure 3 shows the dipoles and mounting system developed. The receiver units combine the linearly polarized signals of the dipoles to produce cir- cular polarized signals, then amplify and bandpass filter these signals, and pass them to the VLA intermediate frequency (IF) system. They also contain an integral noise calibration source. In order to produce serviceable receivers on a short timescale and at low cost, they were constructed almost entirely from commercial components. The VLA signal transmission system allows for two pairs of two parallel-hand signals to be transmitted. The receiver system is designed so that the two senses of circular polarization from the 74 MHz receivers occupy one pair of signal transmission channels while the two senses of circular polarization at 330 MHz occupy the other pair of signal transmission channels. Thus, 74 and 330 MHz observations can be acquired simultaneously. Kassim et al. (1993) used this simultaneous, dual-frequency capability for ionospheric calibration via phase transfer from 330 to 74 MHz (see §4). An alternate signal transmission approach is to use one pair of signals for the upper half of the 1.5 MHz bandpass and the other pair for the lower half, thereby obtaining higher spectral resolution, primarily for radio frequency interference (RFI) excision purposes (§4.2). Measurements of circular polarization are normally available and have been used for both astronomical (solar, T. Bastian, private communication, 1998) – 8 – and RFI excision purposes. 1 Figure 4 presents a complete block diagram of a receiver. In detail, the receiver units are comprised of the following components: 1. The two orthogonal linear feeds are converted by a full quad hybrid into right and left circular polarizations (RCP, LCP). 2. An onboard source injects a noise calibration signal into both the RCP and LCP chains via directional couplers. This source is directly powered by a “Cal” signal provided by the site. 3. Out-of-band rejection filtering is provided by high-Q cavity filters with a center fre- quency of 73.9 MHz and a 1.7 MHz bandwidth. Such a narrow bandwidth is necessi- tated by the close proximity of local television stations. (Tests have been conducted with a 3 MHz bandwidth; the TV signals saturated the receivers.) 4. For the VLA, the signal is transferred directly to the IF system while for the PT receiver, the signal is upcoverted in order to pass into the VLBA IF system, by mixing the RCP and LCP channels with a reference local oscillator (LO) signal. 5. Two power combiners consolidate the 74 and 330 MHz channels for transport to the correlator. The initial work on the 74 MHz system focused on the VLA alone. Concurrently NRAO was in the process of testing fiber optic transmission techniques by tying in the nearby Pie Town antenna (PT) of the Very Long Baseline Array (VLBA). First fringes between a VLA subarray and the PT antenna were observed in 1998 December, and a successful test with the PT antenna and the full VLA was conducted in 1999 September. Routine observations on this facility began in 2000 October. Consequences of this fiber-optic link to Pie Town (the PT link) are that the longest VLA baselines are extended to approximately 70 km, but the number of antennas in the VLA decreases to 26 because the PT antenna replaces one of 1VLA correlator modes ’PA’ or ’PB’ allow obtaining full polarization information at 74 MHz alone, at the expense of halving the number of channels, and thus reducing the ability to purge narrowband RFI. Tests observing a strong, unpolarized source indicate cross-polarization leakage of at least ∼30 %, and since initial attempts at polarization calibration using AIPS task POLCAL failed, users have not been encouraged to utilize this mode. It is possible that a dedicated effort with the EVLA correlator might permit full polarization astronomical measurements in the future, although it is suspected that many, except relatively nearby sources, might be depolarized at this frequency due to Faraday rotation. – 9 – the antennas in the VLA and the VLA signal electronics requires removing another antenna from the array to accommodate the PT signal. Initial efforts on the PT link focused on frequencies above 1000 MHz. The VLBA does have a standard observing frequency at 330 MHz, but it does not have a 74 MHz operating capability. In 2000, NRL and NRAO initiated a program to add a 74 MHz capability to the PT antenna and to operate the PT link at both 74 and 330 MHz. Initial tests were conducted with the 74 MHz receiver replacing the 330 MHz receiver, but the electronics path has been modified subsequently to operate at both 74 and 330 MHz simultaneously. Figure 5 shows the first successful fringes at 74 MHz, obtained on the quasar 3C 123 in the fall of 2001. In the remainder of this paper we shall focus on the VLA system alone. Many of the techniques we describe are equally applicable to observations with the PT link, though experience with that system is considerably more limited and has only been used to observe relatively bright, isolated ob jects (Gizani et al. 2005; Lazio et al. 2006; Lane et al. 2006; Delaney et al. 2004). See Appendix A for an example of the full synthesis VLA+PT image of Cas A (Figure 21). Table 1 summarizes the performance characteristics of the VLA’s 74 MHz system. We quote a sky-noise dominated system temperature that is appropriate for the Galactic polar caps. While the sky-noise for fields on the Galactic plane can be up to ten times higher, in practice the low forward gain of the primary beam smooths out and lowers the variations in Tsys to typically a factor of two or less (§4.3) The sensitivity listed is typical for the A and B configurations, for regions away from the Galactic plane. Sensitivities for the smaller configurations are considerably poorer, because of confusion (§5.1) and presumably because of low-level, broad-band RFI from the antennas and equipment located at the VLA site. Figure 6 shows the primary beam power pattern measured for one of the antennas. Other antennas show similar power patterns. Table 1 cites the primary beamwidth as 11.◦7, but it is clear that the beam has only a modest forward gain and a broad plateau with a poor sidelobe structure. These result from the aforementioned aspects of the system that the antenna itself is only approximately 6λ in diameter and the antenna quadrupod structure. Notice that the sharp edge seen on the right side of the beams presented in Figure 6 is due to the fact that this portion of the beam was not sampled. This was due to motion limitations of the VLA antennas during the holography measurements used to determine the beam power pattern. Nevertheless, this sampling issue did not affect the mapping of the most important part of the beam, down to 20-25 dB from the peak. The poor primary beam definition gives rise to significant sidelobe confusion, as discussed below and further in §5.1. Figures 7 and 8 show the system sensitivity, as measured by the rms noise level in an image, as a function of increasing receiver bandwidth and integration time. This behaviour is typical of observations obtained in the A and B configurations. These images were excised of – 10 – narrow-band RFI, calibrated and imaged following the procedures described in Sections 4 and 5. The deviation from a ∆ν −1/2 dependence with increasing receiver bandwidth in Figure 7 indicates that the system is not thermal noise limited. On the other hand, Figure 8 shows an approximate t−1/2 dependence normally indicative of a thermal noise limited response. Taken together we conclude that we are mainly sidelobe confusion limited since its effects should be independent of bandwidth to first order, but scale roughly as t−1/2 because sources moving through the sidelobes contribute noise in a random walk fashion that averages out with time 2 For integrations of ≥1 hr, the deviation from thermal noise is typically a factor of 2-4. The residual effects of incompletely removed RFI must also play a role in the reduced sensitivity. In fact the sensitivity in the C and D configurations is much poorer than expected from confusion alone, and our hypothesis is that the effects of low-level, broad-band RFI are responsible for that. In practice the 74 MHz system is rarely used in either the C or D configurations. We note that while Table 1 quotes an 8 hour sensitivity limit of 25 mJy, in practice the achievable sensitivity may vary considerably due to the positionally dependent sky noise dominated system temperature, and more importantly from the relative proximity to the handful of extremely bright sources that often dominate the sidelobe confusion (e.g. Cyg A, Cas A, etc.) In §5.1 we discuss both sidelobe and classical confusion further, and show that in the more compact configurations the latter effect can become significant. 4. Calibration of 74 MHz VLA Data Observations of the brightest sources in the sky (e.g., Cyg A, Tau A) with the prototype 74 MHz system demonstrated that the methodologies and algorithms that had been devel- oped for calibration at the standard VLA frequencies were generally sufficient for 74 MHz (Kassim et al. 1993), with self-calibration being particularly important. The larger number of antennas now available makes self-calibration even more robust, but it has also revealed its limits more clearly. In this section we motivate and describe the procedures we have developed for the calibration and imaging of 74 MHz data. The procedure can be summarized as follows: 2One might expect this dependence to disappear once the u-v coverage repeats. In practice, however, position shifts caused by the ionosphere (§4.5) act to “thermalize” the sidelobe confusion contribution. Thus, we believe that the noise due to sidelobe confusion will continue to decrease, at least initially, as t−1/2 even after the u-v coverage repeats. – 11 – Bandpass calibration (§4.1) Observations at 74 MHz are acquired in a spectral line mode, both to enable RFI excision and to avoid bandwidth smearing over the rela- tively large fields of view. Therefore we need to apply a baseline correction to the data. RFI excision (§4.2) Largely because of self-interference, 74 MHz data always must be edited to remove RFI. Amplitude calibration (§4.3) As at higher frequencies, a source whose flux density is presumed to be known must be observed to set the flux density scale. The primary flux density calibrator for the VLA at 74 MHz is Cyg A. Other sources that can be used when Cyg A is not available are Cas A, Vir A, 3C 123 and Tau A. Phase calibration The dominant source of phase corruption at low frequencies is due to the Earth’s ionosphere. Unlike at higher frequencies, one cannot employ as a phase calibrator a source nearby in the sky to one’s target source or field. Instead we have developed two strategies: • When the isoplanatic patch (scale over which the rms phase difference between two lines of sight is approximately 1 radian) size is larger than the field of view of interest (§4.4), a single phase calibration can be applied to the entire field of view. It is most useful in the more compact configurations (C and D) or in the larger configurations (A and B) when imaging a strong source. This strategy relies heavily on current implementations of self-calibration, and it is similar to the calibration strategy used by Kassim et al. (1993). As such, this strategy is a confirmation of their prediction that self-calibration can be used to remove ionospheric phase fluctuations. • When the isoplanatic patch size is smaller than the field of view of interest (§4.5), an angular dependence within the field of view of interest must be used in the phase calibration. We have used a method called “field-based” calibration to do this, which models the ionosphere as a phase-delay screen and uses a grid of background sources to solve for the ionospheric refraction, both globally and differentially within the field of view. Following sections describe imaging requirements (§5) and present examples designed to il- lustrate the efficacy of both phase calibration strategies (§A). This section parallels closely the discussion in our online tutorial3 , which contains more detailed descriptions of the pro- cedures, as well as sample inputs for reducing data within aips. 3http://rsd-www.nrl.navy.mil/7210/7213/LWA/tutorial – 12 – 4.1. Bandpass Calibration As described below (§4.2), the VLA generates considerable internal radio frequency interference. Consequently, observations are performed in a spectral-line mode, which also avoids bandwidth smearing over the large regions (typically) imaged (§5). Characteristic spectral channel bandwidths are 12 kHz in a 1.5 MHz total bandpass. As with any spectral line observation, the amplitude variations across the band (and phase gradients due to delay errors) must be removed by bandpass calibration. Fortunately, the flux density of Cyg A (≃ 17 kJy) is nearly always much greater than the equivalent flux density of any RFI, even in the narrowest channels that the VLA correlator can produce (12 kHz), meaning that one can use observations of it to calibrate the bandpass prior to excising RFI. Other sources—such as Cas A, Vir A, Tau A—can also be used; their lower signal-to-noise ratios on longer baselines can require judicious choices of time or frequency ranges or both in order to calibrate the bandpass. Bandpass calibration is traditionally performed with the assumption that the flux den- sity of the calibrator can be represented by a single value across the bandpass. In the case of 74 MHz observations with the VLA, however, the fractional bandwidth of the system is large enough that the intrinsic visibility of Cyg A can change across it. A rough estimate shows that the effect of resolution can be ignored if π (1) ∆ν ν θsrc θHPBW ≪ 1, where ∆ν /ν is the fractional bandwidth and θsrc/θHPBW is the source size expressed in units of the synthesized beam. In the A configuration, the left-hand side of equation (1) is 0.3, so there is a 30% change (worst case) in visibility across the bandpass on the longest baseline; in the smaller configurations, the error introduced by the fractional bandwidth is proportionally less. Two options exist for dealing with the effects of the intrinsic visibility change across the bandpass. First, one can set an upper limit to the baseline length to be considered, effectively decreasing θsrc/θHPBW . Second, one can divide the observed visibilities at each frequency by a model of the source, thereby transforming the visibilities into what would have been obtained had a point source been observed. In practice, the second option is used most often in order to employ the maximum amount of the data possible, and models4 of Cyg A, Cas A, Vir A, 3C 123 and Tau A are available online. This second method, dividing the visibilities by a model of the source, is done before the 4http://rsd-www.nrl.navy.mil/7210/7213/LWA/tutorial/ – 13 – visibility amplitudes are calibrated. At higher frequencies, the standard bandpass calibration procedure at the VLA is to divide by the “Channel 0” continuum data, formed by averaging the inner 75% of the bandpass. Rather than using a “Channel 0” continuum, which would be contaminated by the presence of the 100 kHz comb (§4.2), one first solves for the individual channel corrections (both amplitude and phase). These corrections are then normalized so that the mean correction across the spectrum is unity. Various weighting schemes can be used during the normalization process. In practice, the weights used in normalization are often taken to be independent of channel; the converse, scaling the weights by some function of the amplitude in each channel, can have the effect of giving more weight to baselines with more RFI (higher amplitudes) and the short baselines (which are most often affected by RFI). In general a single bandpass correction is determined from all observations of the band- pass calibrator, and this correction is applied to all other observations within that “observing run.” For bright sources, residual bandpass calibration errors can introduce systematic er- rors in the image (appearing in the shape of the beam, but which cannot be cleaned) and thereby limit the dynamic range of the image. In these cases, one of two strategies can be adopted. If one has multiple observations of the bandpass calibrator, one can attempt to form the bandpass correction as a function of time. Alternately, during self-calibration one can produce a different phase correction for each channel (or small number of channels). 4.2. RFI Excision Radio frequency interference (RFI) is a significant problem for low frequency VLA observations, though we note that there is little external RFI within the passband of 73.0– 74.6 MHz. RFI is the ma jor limiting factor for the dynamic range of the data observed with the more compact VLA configurations (C and D), due to the interference between the individual antennas of the array. It also constitutes an important issue for the more extended configurations (A and B), and needs to be dealt with appropriately. Essentially all non-astronomical signals are generated internally by the electronics of the VLA’s antennas and (to a lesser degree) signals emanating from other sources on the VLA site. By far the most common is a 100 kHz “comb” generated by the VLA’s monitor and control system. The oscillators responsible for this are located in every antenna. The oscillators can be coherent, with low phase rates such that phase differentials of less than 1 radian are obtained in an integration time, so that the coherence is maintained over long periods of time. The result is spurious correlation, especially between certain pairs of antennas whose oscillators appear to maintain coherence regardless of where they are located in the array. The net (averaged – 14 – over frequency) spurious visibility is not great (. 100 Jy) and can be effectively removed in the spectral domain. (By contrast, RFI at 330 MHz is mainly externally generated but can be excised using the same procedures as described below.) Figure 9 (left panel) shows an example of the 100 kHz “comb” for a single baseline. Various procedures exist to accomplish RFI excision, depending upon the dynamic range requirements. Notice that the final image dynamic range will also depend on the position of the source in the field, usually being lower closer to brighter sources, and how well ionospheric effects can be solved (Sections 4.4 and 4.5). RFI excision appropriate for moderate dynamic range images (DR∼ 103) can be accomplished with the following procedure.5 One begins by identifying a “baseline” region within the spectrum to which only the astronomical source(s) and the noise within the system (i.e., Tsys ) contribute. Usually, this “baseline” region is non- contiguous and does not include the “comb.” Next, visibilities with excessive amplitudes (e.g., 100σ) are flagged, regardless of where they appear in the spectrum (inside the baseline region or not). A linear fit to the “baseline” region is made and subtracted. The ob jective of this linear fit is to avoid (or reduce) the extent to which the flagging is biased by the astronomical source(s) and the system. Finally, visibilities whose residual values exceed a user-defined threshold (typically 6–8σ) are flagged in channels both inside and outside of the “baseline” region. The results of this procedure are presented in the right panel of Figure 9, where we show the example of the visibility spectrum on a baseline after RFI editing. An inspection of this image shows that this procedure has eliminated the RFI emission present in the left panel. A similar procedure, also appropriate for moderate dynamic range imaging, is to flag all of the channels comprising the 100 kHz comb6 and then clip any remaining data with excessively large amplitudes. While elementary, these approaches have proven to be reasonably effective in removing RFI at 74 (and 330) MHz. Additional tests also have been found to be useful. For instance, channels in which the ratio of the real part of the visibility to the imaginary part exceeds a given value can be flagged. Various observers have noted that RFI can often be strongly circularly polarized so a test on the amplitude of the circular polarization (Stokes V param- eter) in each channel is also useful. For the highest dynamic range images (DR∼ 104), it is often necessary to inspect the data by hand and perform additional editing.7 Lane et al. 5Within aips, this procedure is implemented using FLGIT or by VLAFM, a special-purpose task added to aips. 6Within aips, the task SPFLG can be used. 7 Within aips, the tasks SPFLG, TVFLG, and WIPER can be used either singly or in combination. – 15 – (2005) describe in more detail post correlation RFI excision with the VLA and VLBA (see also Golap et al. (2006)). The limited spectral resolution of the VLA’s original correlator means that even the narrowest channels provided are 12 kHz wide. As a result, the procedure we describe can remove a significant fraction of the data—typically 10 to 25%. A key design goal of any future correlator would be to provide considerably more channels with higher frequency resolution, allowing much more precise removal of these artificial signals. Such improvements are forthcoming for the 74 MHz system with the transition to the EVLA and its WIDAR correlator. Furthermore, techniques of pre-correlation RFI excision would clearly offer key advances over current approaches. A final note is that the discussion above focuses on narrow-band RFI, much of whose source is relatively well understood. Since the sensitivity and dynamic range in the rarely used C and D configurations appears worse than attributable to confusion or poorly excised narrow-band RFI, we hypothesize that an additional form of low-level, broad-band RFI is limiting the system performance in those configurations. That suggests that the source of interference is the VLA itself, or equipment at the site. Clearly another lesson for future instruments is design them with a focus on eliminating or at least shielding against self- generated RFI. 4.3. Amplitude Calibration 4.3.1. Low-Frequency Calibration Problems and Advantages of the VLA System Despite relatively limited sensitivity compared to that typical of higher frequency obser- vations, the sensitivity of the 74 MHz VLA is unprecedented for frequencies below 100 MHz, and so it has now become possible to obtain flux density measurements for thousands of radio sources that have never been measured before at these frequencies. However, it is important to note that flux density scales are rather uncertain below 100 MHz for several reasons. First, the Galactic background dominates system noise temperatures, so system temper- atures vary as an antenna is scanned across the sky and system gain and temperature must be measured continuously. Antenna gains vary at less than the few % level as a function of time and are consistent with elevation dependent deformation of the feed or dish structure being negligible at 4-m wavelength. This trend is consistent with elevation-dependent gain variations at L-band (20-cm) measured at less than 1%, while variations at P band (90-cm) have never been seen. Tsys certainly changes with time, as it depends on Tsky which is a strong – 16 – function of sky position, being much greater on the Galactic plane than off. Early single antenna tests showed variations of up to a factor of 2 between “on” and “off ” regions against the Galactic plane - being greatest towards the Galactic center. This is much less than the known contrast in brightness temperature between inner regions of the Galaxy and cooler parts of the extragalactic sky as determined by early, lower resolution measurements (e.g. Yates 1968). However that real variation is diluted by the poor directivity of the 25-m dishes at 74 MHz, with ≤20% of the power in the main beam and the rest scattered in sidelobes across the sky. Hence while Tsys does change as a function of time, the variation is greatly smoothed out and is normally less than a factor of two over the course of any pointed, full synthesis observations. A related consideration is that calibrator observations of calibrators like Cyg A not drive the system response into a regime in which a non-linear correction of its correlation coefficient is required. Fortunately the increase on Tsys when observing Cyg A is modest. A rough estimate from existing low resolution sky maps suggests an increase in Tsys of ≤40% in the main beam alone, again consistent with early single antenna tests. Thus the contribution from the distributed “hot sky” still dominates the measured power, and are consistent with the measurements described below indicating that variations in system gain and temperature are being tracked accurately (see also Sections 3 and 5.1). Second, many systems utilize fixed antennas whose power patterns are not very well de- termined. The relative flux densities of sources at adjacent declinations can be determined reliably because they traverse the same parts of the fixed antenna’s pattern, but measure- ments of sources at widely separated declinations are difficult. Also, many measurements must be made far from the maxima of the primary response patterns of the antennas, where the patterns are less stable than at the maxima. Finally, ionospheric amplitude scintillations often disturb measurements below 50 MHz and ionospheric absorption can affect them below 20 MHz. The VLA has features that avoid most of these problems. First, the gain and noise temperature of each antenna are monitored continuously by the injection of noise calibration signals into every preamplifier. At higher frequencies, experience has shown that the gain properties of the VLA electronics system are stable at the 0.1% level, and that linearity is preserved even if the Tsys changes by a factor of 3 or more. Second, the VLA antennas are pointed at each source under observation, and the antenna power pattern is stable, at least for measurements made near its center. Sources at different declinations should be directly comparable. Third, 74 MHz is a high enough frequency that ionospheric amplitude scintillations and absorption are minimal, though occasional episodes of scintillations have been seen during observations of strong sources. – 17 – 4.3.2. VLA Calibration Method Because of the high system temperature of the 74 MHz system, the best source to use for calibrating the visibility amplitudes is Cyg A. However, as it is partially resolved, a good model is required for accurate amplitude and phase calibration. We have used multiple observations in multiple VLA configurations to generate such a model, and it is available online8 . We base our flux density scale on that of Baars et al. (1977), in which Cyg A has a flux density of 17 086 Jy at 74 MHz. Although Cyg A is by far the best amplitude calibrator, other sources can be employed, if necessary. Vir A is acceptable in all VLA configurations, while Tau A and Cas A can be used in the smaller configurations. As with Cyg A, all of these sources are resolved significantly, so model images (available online) are necessary for calibration. The amplitude calibration is done in the usual way in AIPS, with the task CALIB used to derive the antenna complex gains. Residual amplitude errors are corrected by self calibration in the final stages of the reduction. As described below, measurements demonstrate that reliable fluxes, at the 5% level, can be obtained down to the half power beam point. In the rare case of severe ionospheric weather conditions such as ionospheric scintillations (§4.5.1), amplitude correction is not possible and the data are discarded. 4.3.3. Reliability of the VLA flux-scale at 74 MHz In order to quantify our ability to cope in practice with the problems described above (§4.3.1), it is necessary to examine the robustness of the flux density scale for the VLA. As part of a snapshot survey (§A), we observed a number of strong sources, which we have used to constrain the gain stability of the 74 MHz system. Kuhr et al. (1981) presented the spectra of 518 extragalactic radio sources that have flux densities above 1 Jy at 5 GHz, using data compiled from many catalogs. The absolute flux density scale was based on that of Baars et al. (1977). Although the data below 100 MHz are rather sparse, the spectra of Kuhr et al. (1981) have proven to be quite reliable in most cases. They do not include any sources near the Galactic plane and give analytic expressions for spectra only when they are fairly simple (straight or moderately curved). Sources with complex spectra are included in the catalog, but no attempt was made to fit such spectra with analytic expressions. Of the 29 sources in our snapshot survey, eleven have spectra for which Kuhr et al. (1981) give analytic expressions. Table 2 compares our measured 74 MHz flux densities with 8http://rsd-www.nrl.navy.mil/7210/7213/lazio/tutorial/ – 18 – the value predicted at 73.8 MHz from those expressions. (See also Table 8.) Because of their relatively simple structure even at the highest angular resolution, the four most reliable sources for comparison appear to be 3C 98, 3C 123, 3C 219, and 3C 274. These yield an average flux density ratio (VLA/Kuhr et al. 1981) of 0.99 ± 0.06. A simple average of all eleven ratios yields 0.98 ± 0.13. In either case the agreement between the Kuhr et al. (1981) and VLA flux density scales appears to be as good as 2%. We conclude that 74 MHz flux density measurements are stable at the few percent level, and that the absolute flux density scale as tied to the Baars et al. (1977) value for Cygnus A is accurate to at least 5%. These results also confirm that antenna system temperatures and gains are being tracked correctly, and that power is being detected linearly. Hence the 74 MHz flux scale is reliable for those sources never observed before at this frequency. However, Kuhr et al. (1981) give correction factors that should be used to adjust the data from the various catalogs to their flux density scale. When making comparisons, these corrections should be used. 4.4. Angle-Independent Phase Calibration The phase calibration procedure employed commonly at centimeter wavelengths makes use of a secondary calibrator, a modestly strong source much closer in the sky to the target source than the amplitude calibrator. It is chosen to be strong enough to dominate the total flux within its field of view, so that a unique phase can be derived for the contribution of the array’s electronics. In contrast, at 74 MHz the field of view is so large that few sources are strong enough to dominate the field. Thus, an initial phase calibration is determined from Cyg A (or another strong source such as those listed above, §4.3), which usually provides sufficient coherence on enough short (≤ 10 km) baselines to form an initial image even if the target source is in a completely different portion of the sky. Oftentimes, coherence is retained even after the phases are smoothed over long time scales, even longer than the duration of the observation. This technique works only for bright sources. Implicit in this procedure is the assumption that a single phase suffices for the entire field of view—hence its angle-invariant designation. This initial phase calibration works because relatively larger scales (& 50 km, i.e., larger than the VLA) present in the ionosphere dominate the phase fluctuations. It is also the reason previous low frequency (< 100 MHz) systems with baselines less than about 5 km were able to function without ionospheric correction techniques—except for a constant refractive shift arising from the large-scale structures, short baselines (. 5 km) were, to first order, – 19 – relatively unaffected by the ionosphere (Erickson 1984)9 . The initial image obtained from this calibration must be subsequently astrometrically corrected and self-calibrated. We write the observed phase on a given baseline as φ(t) = φsrc(t) + φVLA + φion (t) (2) where φsrc is the phase contributed by the intrinsic structure of the source, φVLA is the phase contributed by the VLA electronics, and φion is the ionospheric contribution. We write φ(t) to emphasize that the phase is a function of time, not only because the Earth-rotation synthesis means that the phase can change as the array samples different portions of the u-v plane (φsrc) but also because of the temporal variations imposed by the ionosphere (φion ).10 During calibration observations of a strong source, the first term φsrc can be calculated and removed using a model of the source, so we shall ignore it henceforth. The smallest scales in the ionosphere over which the rms phase varies by more than 1 radian at 74 MHz are typically no smaller than 5 km. As a result, the ionospheric phase term, φion , can be considered to be relatively constant across at least the short baselines in the array (. 5 km). Transferring the phases determined toward the calibrator source (i.e., Cyg A) to another direction in the sky introduces a term, φion,cal , the ionospheric phase distortion toward the calibrator. However, for the short baselines, this term is effectively constant. It introduces little more than a refractive shift of the apparent source positions (Erickson 1984). This refractive shift can be removed by registering the image with an existing all-sky survey at a higher frequency (i.e., the NVSS, Condon et al. 1998; WENSS, Rengelink et al. 1997; and SUMSS, Bock et al. 1999). This initial, crude phase calibration is sufficient to produce an image. The initial image then serves as the initial model for hybrid mapping, which consists of iterative loops between self-calibration and imaging. With the large number of antennas in the full VLA, convergence occurs rapidly. In this respect, the process is similar to that employed often in very long baseline interferometry (VLBI) in which a crude initial model is combined with several iterations of hybrid mapping to produce the final image (Walker 1995). Phase self-calibration is warranted only if a sufficient signal-to-noise level can be ob- 9Short baselines are undoubtedly affected by small scale ionospheric structure beyond simple refractive shifts, but those effects were negligible for past arrays where the intrinsic sensitivity was already confusion limited to Jy-levels. As a new generation of low frequency arrays aims to achieve much greater levels of sensitivity, the phase variations due to small scale ionospheric structure may well be the limiting factor. 10 In practice, φVLA can also be a function of time because of phase jumps in the electronics. These are sufficiently rare that we shall ignore them. – 20 – tained (Cornwell & Fomalont 1999). In practice there are two signal-to-noise level thresholds that must be met, but both can be exceeded quite easily. The first signal-to-noise threshold that must be met is that there must be a source or sources that are strong enough to be detected in the approximate ionospheric coherence time (≈ 1 min.). Both an extrapolation of higher frequency source counts (Bridle 1999) and source counts derived from 74 MHz ob- servations indicate that there should be roughly 150 Jy of flux from different sources within the primary beam, originating from sources stronger than about 5 Jy, which is about 5 times the rms noise level obtained in a 1 min. integration time. The second signal-to-noise threshold is that the phase derived from the calibrator source must dominate over weaker sources in the field. This criterion can be understood by consid- ering the calibrator, with a flux density Scal , to be immersed in a “sea” of randomly-located background sources, with a typical flux density Sb . Treating the visibilities of the sources as phasors, the background sources will contribute to a jitter of the phase determined for the calibrator. A rough estimate of the magnitude of this jitter is √N (3) δφ ∼ Sb Scal , where N is the number of background sources in the field of view. For the purposes of a rough estimate, we take δφ ∼ 0.2 (implying a phase jitter signal-to-noise threshold of 5). With Sb ∼ 5 Jy and N ∼ 40(= 200 Jy/5 Jy), we find Scal ∼ 150 Jy. Though clearly a rough estimate, experience has shown that it is possible to achieve successful phase self-calibration with as little as 50 Jy from bright sources in the initial model. More generally, nearly every randomly-picked field of view will contain at least one 3C ob ject (or equivalent at southern declinations) whose flux density alone is close to this minimum value, and existing all-sky surveys at higher frequencies can be used to identify the (few) strongest ob jects totaling at least 100 Jy within the field of view. (The phase fluctuations in the top panel of Figure 5 are consistent with the estimate derived here for the phase jitter due to background sources.) A potential weakness of this method is its limited utility to the larger configurations (A and B). Depending upon the state of the ionosphere, it may be difficult or impossible to obtain sufficient signal-to-noise on the more limited number of short baselines. In turn, this may impair one’s ability to produce an initial model for self-calibration. An alternate phase calibration strategy does not rely on the phases transferred from a distant source. Instead, a strong source, like Cyg A, is used to establish the amplitude scale. An all-sky survey (particularly the NVSS or WENSS) is used to construct a sky model11 11Within aips, the sky model is constructed using FACES. – 21 – within the primary beam; for sources north of declination +30◦ , the lower frequency WENSS is preferable to NVSS. This sky model then serves as an initial model for self-calibration. In effect, no attempt is made to use an external calibrator to calibrate the phases. A primary benefit of this approach is to produce a map with astrometrically correct positions, as the coordinate system of the map is locked to the a priori known position of sources in the sky, while self-calibration with a model produced from raw data locks the final position to an arbitrary sky position determined by the position of the model or ionospheric refraction. This alternate strategy is a simplified application of the field based phase calibration technique, described in the next section. 4.5. Field-Based Phase Calibration The prior calibration strategy is most useful at 74 MHz within a restricted field of view containing a strong source or strong sources located relatively close together. In this section we first motivate why phase angle independent self-calibration is insufficient for all observa- tions, then describe the method we have developed to handle more general observations. 4.5.1. Motivation The isoplanatic patch is the characteristic scale over which the rms phase difference between two lines of sight is approximately 1 radian. At low frequencies the size of the isoplanatic patch is determined primarily by the ionosphere. The phase contributed by a cold plasma is φ = reλ Z D 0 = reλNe , ds ne (s), (4) where re is the classical electron radius, λ is the free-space wavelength, ne is the electron number density, and Ne is the electron column density (also known as the total electron column, TEC). For reference, two lines of sight for which ∆φ ∼ 1 rad at 74 MHz would differ in electron column density (TEC) by ∆Ne ∼ 1014 m−2 . Two lines of sight separated by the diameter of the primary beam pierce the F-layer of the ionosphere (the most dense region of the ionisphere at an altitude of ∼400 km) at a linear separation as large as 80 km. This is larger than the size of the array itself, even with the PT link, and the ionosphere can have significant column density variations at much smaller scales (. 10 km), so the primary beam typically includes multiple isoplanatic – 22 – patches. Figure 10, which itself is a portion of a larger image, shows the effect of constructing an image larger than the isoplanatic patch. Clearly evident is a systematic distortion of the sources attributable to incorrect phase calibration. Figures 11–14 illustrate the successively higher order ionospheric effects that lead to the break-down of a simple angle-invariant self-calibration. Figures 11 and 12 shows the effects of the largest scale (> 1000 km) ionospheric structure, a “wedge” that acts to shift the entire field of view, without source distortion, on time scales of minutes. This structure dominates the total electron content (TEC) and also causes Faraday rotation of linear polarization. A GPS-based method to correct this effect has been demonstrated at the VLA (Erickson et al. 2001), but self-calibration alone can compensate for it if the time scale on which the phase corrections are calculated is sufficiently short to track this gross refraction. It will, however, leave the field with a gross astrometric offset from the correct source positions. Figure 13 shows the phase effects imposed by ionospheric mesoscale structure. These mesoscale structures are due typically to traveling ionospheric disturbances (TIDs), with scale sizes on the order of hundreds of kilometers, column densities of order 1015.5 m−2 and periods less than 90 minutes. While the size of the VLA—even in its A configuration—is smaller than that of a TID, it is still sufficiently large that TIDs can impose a (mainly) linear phase gradient down the arms of the VLA. The effect of this linear phase gradient can be seen as a (nearly) linear increase in the phase with increasing distance from the array center; the amplitude of the offset also increases with baseline, as the antenna separation becomes a larger fraction of the TID. (In the example shown in Figure 13, the traveling nature of the disturbance is also present as a lag between the times when the maximum phase offset is obtained at the various antennas.) Self-calibration can remove the effects of TIDs only for a limited region in the field of view. At any given instant, sources outside this region are “differentially refracted” or shifted by varying amounts in proportion to their distance from nominal direction of the self-calibration solution. Because a typical observation lasts much longer than the time it takes a TID to pass over the array, the differential refraction changes with time. Thus, sources in a map made using all of the data are blurred as well. Jacobson & Erickson (1992a,b) conducted an extensive study of acoustic gravity wave generated TIDs having phase speeds less than 200 m s−1 and periods greater than 103 s using the VLA at 330 MHz. They found them to have a quasi-isotropic distribution in azimuth. On shorter time scales (< 300 s) they found the ionospheric phase effects to be dominated by faster (> 200 m s−1 ) magnetic eastward-directed disturbances (MEDs). Over the baseline relevant to the extended VLA configurations and the PT link (∼ 50 km), both TIDs and MEDs contribute to 74 MHz phase effects. Figure 14a-f were obtained from an 8 hour observation towards a strong source (Virgo – 23 – A), and illustrate a variety of effects due to ionospheric phenomena typically observed at the VLA on scales sizes both larger and smaller than TIDs. Figures 14 a and b illustrate first order effects on the visibility phase towards Virgo A. Figures 14 c and d track the flux density and apparent position of Virgo A, while Figures 14 e and f reflect differential refraction towards five field ob jects after the first order term has been removed. These differential phase effects due to smaller scale ionospheric structures, on scales of tens of kilometers, are the most intractable and pose A MAJOR challenge for future instruments. (The high frequency “jitter” superposed on the TID-produced structure in Figure 13 is also contribution from smaller scale structure.) These phase effects are generated by turbulent structures comparable to OR SMALLER than the size of the VLA, and lead to source distortion. There is no current means of correcting for these effects over a wide field of view and when they are severe, the data must be discarded. If these small-scale structures generate phase distortions larger than 1 radian on scales smaller than the size of the VLA antennas, the result is ionospheric scintillations (also illustrated in Figure 14). We believe that there will never be a means for compensating for this effect, at least so far as imaging applications are concerned, as the data are effectively smeared in phase before they arrive at the antennas. (Ionospheric scintillations can be useful for studying the ionosphere itself, though.) Fortunately ionospheric scintillations are rare at the VLA at 74 MHz, but when they do occur those portions of the data must be removed. In principle, a data-adaptive calibration scheme, based on self-calibration, could be used to remove the phase errors that result from imaging a region larger than the isoplanatic patch. Current implementations of self-calibration12 model the phase error on the ith antenna at time t as a single quantity, δφi (t), equivalent to assuming that the isoplanatic patch has an infinite extent above the array. The key assumption for self-calibration—that errors can be modeled as being antenna-dependent—remains true for a data-adaptive scheme. Thus, a joint multi-source self-calibration, in which the calibration correction would become a function of position on the sky, δφ(t; α, δ), seems possible. Limited testing suggests that such a scheme could work, though it probably requires considerably better signal-to-noise than can be achieved with the VLA; no comprehensive attempt to implement and assess such a scheme has been performed yet. In principle such a technique might allow wide- field imaging across the full extent of the primary beam on the longest baselines allowed by natural limits of brightness temperature sensitivity (for a related discussion, see also Erickson (2006)). When it is desired to image an entire field of view, the assumption of an infinite isopla- 12This description includes CALIB within aips. – 24 – natic patch is no longer valid. The effects of such an assumption can be seen as a tendency to detect a larger number of sources toward the direction of the strongest source in the field (Cohen et al. 2003). Figure 15 illustrates this effect in a field containing 3C 63 (≈ 35 Jy). The density of sources across the field is clearly non-uniform. Besides the effects of anisopla- naticity, which shows up at scales of 5-10◦ , the non-uniform density can also arise from two other related causes. First, if sufficiently high-order ionospheric distortions are present, they contribute to increasing phase errors at increasing distances from the effective phase center. Sources at large distances will be blurred, thereby decreasing their brightnesses. Second, Figure 16 demonstrates that the apparent source positions also wander over time. Because the amount and direction of the wander can vary both over the field of view and with time, sources are smeared further and their brightnesses decrease further. Only in the neighbor- hood of a strong source are the self-calibration solutions able to track the ionospheric phase distortions accurately. If one is interested in imaging only a strong source, the non-uniform distribution of sources is usually unimportant. If a field does not contain a strong source, the self-calibration solutions represent some “average” ionospheric phase distortion across the field. Again, depending upon the scientific problem being attacked, using angle-independent self-calibration may be sufficient. 4.5.2. Implementation Compensation for higher order ionospheric phase distortions is required (Cotton & Condon 2002; Cotton et al. 2004), when the entire field of view is of interest (e.g., in the VLA Low- frequency Sky Survey Cohen et al. (2003, 2006)) or when the ob ject of interest is “close” to a strong source. The latter case happens fairly frequently given that the typical separation of 3C sources on the sky is approximately 8◦ , comparable to the size of the VLA field of view. Moreover, the dynamic range required to detect or image a weak source close to a strong source may impose much more stringent constraints on the need to determine the ionospheric phase than the approximate 1 radian criterion that we gave at the beginning of the previous section. In such cases, we model the ionosphere as a phase-delay screen.13 We return to equa- tion (2), assume that φsrc is calculable and can be removed, and expand the ionosphere term as φ = φVLA + φion,lo + φion,hi . (5) 13Within aips, this procedure is implemented using VLAFM, a special-purpose task added to aips. – 25 – Here φion,lo refers to the phase distortion introduced by low spatial frequency structures in the ionosphere while φion,hi refers to high spatial frequency structures. In order to make a division between “high” and “low” spatial frequencies (or “long” and “short” wavelengths) we make a “frozen-flow” approximation in which the ionospheric structures are assumed not to change internally over the time it takes for them to be transported across the array. Under this assumption, spatial and temporal scales become equivalent. The smallest structures (. 10 km) are transported at speeds of order 100 m s−1 on time scales of order 100 s. We consider these to be high spatial frequencies. Structures of 100 km or larger are transported across in time scales of order 15 min.; these are low spatial frequencies. The large-scale structures are larger than even the maximum baselines of the A config- uration. We therefore decompose the low-frequency ionosphere term into low-order Zernike polynomials nZ l Al n (6) φion,lo = 2 Xn=1 Xl where Z l n is the Zernike polynomial, Al n is the coefficient of that polynomial, and the standard conditions apply that n ≥ l and n − l is even. The n = 1 terms account for the large-scale refractive shift of the field of view while the n = 2 terms describe astigmatism or differential refraction within the field of view. The n = 0 term is not used because it represents an overall phase advance or delay (“piston”) to which the interferometer is insensitive. We use Zernike polynomials because they represent a class of polynomials orthogonal on a circle. Thus, they are useful for representing distortions across the aperture of the array.14 Table 3 summarizes the polynomials used. The methodology is quite similar to that used in adaptive optics systems in optical astronomy. One important difference between our use of the Zernike polynomials and that in adaptive optics, however, is that these polynomials are used to describe wavefront errors in or near the aperture for the case of adaptive optics systems in optical astronomy whereas here they represent errors quite far from the aperture plane. In order to derive the required corrections, snapshot images of sources in an “astrometric grid” are produced, and the offsets between the apparent and expected locations of sources in the astrometric grid are determined. The snapshots must be formed on short enough timescales so as to track the ionospheric phase variations, typically 1 min. or shorter. Both the NVSS and WENSS can be used to produce this astrometric grid as both are constructed at a high enough frequency and from a sufficiently large number of sources that the positions 14The phase-delay screen representation and Zernike modeling of the ionosphere is not strictly correct as the ionosphere is not in the far-field of the array. Nonetheless, we view the Zernike polynomials as a useful first step in modeling the ionosphere. – 26 – of sources within these catalogs is known to an accuracy much better than the synthesized beam at 74 MHz, even in the A configuration. From the source offsets, the coefficients of the Zernike polynomials can be found in a least-squares minimization. Figure 17 shows an example of a (subset) of an astrometric grid and the resulting φion,lo over the VLA for a particular observation. The snapshot images of the astrometric grid sources are useful only if two conditions are met. First, the ionosphere must be stable enough that sources are not defocused seriously but merely shifted from their expected positions by refraction. Second, one must be able to determine φVLA prior to forming the snapshot images. As in §4.4, the initial phase calibration is determined from observations of Cyg A, or another primary calibrator. For a single source at a well-known position, one need solve for only the n = 1 terms describing an overall refractive shift. Correcting for this global refractive shift should yield φVLA . If φion,hi is non-negligible (meaning that sources may be defocused), it will corrupt estimates of φVLA . Because no phase calibration is performed during this strategy (but see below), any errors in determining φVLA remain in the data during all subsequent processing. Figure 18 presents the same field as in Figure 15, only this time calibrated using this field- calibration strategy. The density of sources across the field is seen to be far more uniform. Figure 19 quantifies the improvement that the field-based calibration; it shows the rms jitter in the apparent separations of pairs of sources of various separations. This rms jitter is a fairly direct measure of the refraction differences induced by the low spatial frequency ionospheric irregularities. For separations greater than 2◦ , the phase screen corrections dramatically reduce the jitter. Also important is that the rms jitter shows no trends as distance from the phase center increases. One weakness in our current implementation of this strategy is that the ionosphere is modeled in a piece-wise fashion in time. No “smoothness” constraint is applied to Zernike models from adjoining snapshots. Work is ongoing to rectify this weakness. In order to assess if (or to what extent) φion,hi is corrupting our estimate of φVLA , we observe Cyg A multiple times during the course of an observing run. By comparing or, more often, averaging the various estimates of φVLA + φion,hi , we seek to minimize the contribution of φion,hi to our estimate of φVLA .15 We close this section with a few general comments on our choice of Zernike polyno- mials and our implementation. From the work of Jacobson & Erickson (1992a,b) showing 15This process is implemented in the task SNFLT, a custom-designed task designed to work within aips and available from the authors upon request. – 27 – that much of the ionospheric structure above the VLA is in the form of waves, one might wonder if a Fourier representation would not be more appropriate. We have chosen to use Zernike polynomials to describe the ionospheric structures precisely because they were invented for the purpose of describing phase errors across a circular aperture. Moreover, compared to a rectangular Fourier transform, many fewer terms of Zernike polynomials are required to describe the ionospheric phase fluctuations. Given the limited sensitivity of the VLA, this criterion is quite important. Finally, although it is not yet contained within our implementation, a natural extension of our method would involve requiring the ionospheric phase corrections to be smooth in time. In order to impose this requirement, one requires an orthogonal basis for the modeling since the interpolation in time is by interpolating the coefficients and this only works if the terms being interpolated are orthogonal. We emphasize that our division between φion,lo and φion,hi is phenomenological and has the effect of making the division based on the order of the Zernike polynomials used rather than on physical properties of the ionosphere. The range of spatial scales in the ionosphere implies that one could use higher-order Zernike polynomials to decompose the phase distor- tions. In principle, by incorporating a sufficient number of Zernike polynomials one could reduce φion,hi to a sufficiently low level so as to be unimportant; orders as high as 80 are not unusual for correcting large optical telescopes. However, the number of Zernike coefficients that can be determined is limited by both the sensitivity of the system and the aperture distribution of the VLA. The sensitivity of the VLA is an issue because the field must be imaged on less than the ionospheric coherence time, which in turn depends upon the array configuration as the amount of phase contributed by the ionosphere can depend upon the baseline length (Figure 13). In 1 min., a coherence time appropriate for the B configuration, a typical field of view contains no more than 15–20 sources strong enough to be detected; for A configuration, a more typical coherence time is approximately 20 s. Even during times of ionospheric quiescence, fewer than 10 sources may actually be detected; during poor iono- spheric “weather” conditions or when sidelobes from strong sources obscure weaker sources, the actual number detected with any confidence may be no more than 5 sources. Hence, in order to avoid (or minimize the occurrence of ) spurious Zernike coefficients, we have re- stricted the modeling to only the n < 3 terms. This model of the ionosphere is constructed anew every 1–2 min. The aperture distribution (as opposed to the more traditional concerns in interferom- etry regarding the u-v distribution) is related to the range of spatial scales sampled in the ionosphere. In order to characterize large-scale ionospheric structures (e.g., TIDs) that are of concern under typical “weather” conditions, only a modest number of pierce points through the ionosphere are required. In this respect the aperture distribution of the compact config- urations (C and D) provides adequate sampling of the relevant spatial scales, though such – 28 – sampling is also often not needed because the array remains nearly coherent in these con- figurations. In contrast, in order to characterize smaller scale ionospheric structures, a high density of closely spaced pierce points is required, which in turn requires high spatial frequen- cies in the aperture plane, rather than in the u-v plane. In the extended configuration (A and B), the sparseness of the aperture distribution means that small spatial scales are hardly sampled at all. The design of future low frequency instruments may require a compromise between the good uv coverage important for imaging and the good aperture plane coverage that might be required to allow sufficient modelling of small scale ionospheric structure. 4.6. Phase Transfer Under especially severe ionospheric weather conditions (with ionospheric phase rates on long baselines in excess of 1 deg s−1 ), it may become necessary to scale the 330 MHz ionospheric-induced phase rates, transfer, and remove them from the 74 MHz data stream (Kassim et al. 1993). In practice, this dual-frequency ionospheric phase transfer technique has been required only rarely, as even in the A configuration phases transferred from a strong source anywhere in the sky retain sufficient coherence on enough short spacings to provide an initial model for self-calibration. Subsequent iterations of self-calibration then improve the phases on the longer baselines. As with straight-forward self-calibration, phase transfer does not compensate for the main failing of self-calibration, which is the lack of an angular dependence on the antenna based phase solutions. However phase transfer may well be of great benefit to future low frequency instruments, especially those planned to operate at lower frequencies and longer baselines than the VLA, such as the LWA (Kassim et al. 2004, 2006b) and LOFAR (Kassim et al. 2004). Therefore it is important that their design does not preclude the possibility of simultaneous observations at multiple frequencies. 4.7. Evolving Techniques of Phase Calibration The previous sections illustrate successive schemes of ionospheric phase calibration that continue to evolve with real observational experience. A technique such as joint, multi- source self-calibration is required to realize the full potential of emerging larger low-frequency instruments such as the LWA and LOFAR but has yet to be developed and tested. Other related appoaches to the calibration of large, low frequency arrays continue to be proposed and discussed (Erickson 2006; Nijboer, Noordam & Yatawatta 2006; Cotton 2006; Brentjens 2005; Noordam 2004; Lane et al. 2004b). – 29 – Table 4 presents a simple overview of these evolving schemes; the illustrations in Cotton et al. (2004) may help orient the reader with respect to the applicable geometries. 5. Imaging at Low Frequencies In this section we summarize the essential procedures needed for effective imaging in the presence of effects that are particularly significant at these low frequencies. 5.1. Confusion and Sidelobes A key limitation to previous low frequency interferometers operating below 100 MHz has been the poor and confusion-limited sensitivity that arose from their low angular resolution. The high angular resolution of the larger (A- and B-) configurations of the VLA provides some mitigation, but the poor forward gain of the antennas increases the confusion. In fact, two effects are at work, and we distinguish between classical confusion and sidelobe confusion. The former occurs when the density of sources within the synthesized beam becomes so large that they cannot be separated (e.g., 1 source per every 10 beams is a common criterion for the onset of classical confusion). The latter results from the incompletely removed sidelobe response to bright sources. Table 5 summarizes the classical confusion limits for the four VLA configurations, where we have taken classical confusion to occur when there is one source per 10 synthesized beams. In order to estimate these classical confusion limits, we have adopted a log N -log S relation derived from existing 74 MHz observations, 1 Jy (cid:19)−1.25 N (> S ) = 1.25 deg−2 (cid:18) S where N (> S ) is the areal density of sources, in units of deg−2 , stronger than S Jy. Strictly, this log N -log S relation is valid only for flux densities S & 0.25 Jy, based on the flux densities of the sources from which it was determined. Expectations of source densities, based on higher frequency source counts, suggest that this log N -log S relation will tend to overestimate the confusion flux density levels, though, and estimates using source counts scaled from higher frequencies give similar results. Even with the uncertainty in the low- frequency log N -log S relation, the results presented in Table 5 show that classical confusion is unlikely to limit the sensitivity of 74 MHz observations in the A configuration and probably not in the B configuration (possibly only in crowded regions like the Galactic center), but it is a serious factor for C- and D-configuration observations. (7) – 30 – Sidelobe confusion results from the improperly subtracted response to bright sources both inside and outside the main field of view, and is exacerbated when those sources are unresolved by the synthesized beam. The dominant effect is often from the pathlogically brightest sources (e.g. Cyg A, Cas A, etc) outside the main field of view whose emission “rumble” through the primary beam sidelobes. This problem is particularly severe for the 74 MHz VLA because of the relatively small aperture (≈ 6λ) and the sidelobes caused by the feed support structure. Thus the primary antenna gain is low, the primary beam is large (FWHM ≈ 11◦), and the sidelobe levels are high and asymmetric (peak sidelobes are typically only 20 dB down from the main lobe), with a large area of sky in the “close-in sidelobe” region (Figure 6). Moreover, the receiver noise is a small fraction of the total system temperature, so ob jects all over the sky, including those seen through the sidelobes, produce measurable coherence. Because of this “signal-rich” environment at 74 MHz, it is often necessary to image the entire primary beam and a small number of sources outside it. Competing factors mitigate the effects of sidelobe confusion, including delay beam and ”ionospheric” smearing of the residuals of subtracted sources. Nevertheless remaining arti- facts from the few, strongest sources outside the main field of view often limit the sensitivity and dynamic range. Sidelobe confusion can still dominate over classical confusion even in the more compact configurations, since in those cases the arc-minute size scale of the nor- mally offending sources (most notably Cyg A, Cas A, Vir A, Her A, Hyd A, and Tau A) prevents them from being resolved out, and hence their residual effects are worse. (The dominance of sidelobe confusion over thermal noise was indicated earlier in Figures 7 and 8 and discussed In §3. As noted earlier, the sensitivity in the compact configurations is worse than attributable to either form of confusion alone, and our hypothesis is that low-level, broad-band RFI, possibly self-generated, are limiting the performance. This situation is in direct contrast to that at centimeter wavelengths (Condon et al. 1998, viz. Figure 15) where the receiver temperatures dominate the sky temperature, a more uniform aperture illumination produces lower sidelobes, and the non-thermal spectra result in most sources being fainter than at low frequencies, so that only rarely does one have to contend with sources outside the primary beam. 5.2. Wide-field Imaging The standard two-dimensional Fourier inversion of visibility data requires that wθ2 ≪ 1 where w is component of the interferometric baseline in the direction of the source and θ is the field of view. This assumption is not valid for 74 MHz observations over any significant – 31 – hour angle.16 Consequently, inverting the visibility data requires either a three-dimensional Fourier inversion or “polyhedral imaging,” in which the field of view is tessellated into facets over which the assumption of a two-dimensional inversion is valid (e.g., Figure 18). For further details and examples, see Cornwell & Perley (1992) and Perley (1999). In general, polyhedral imaging is the more common technique.17 Table 6 gives general guidelines as to the size and number of facets required for acceptable polyhedral imaging over the entire primary beam. In constructing Table 6 we have made use of the criterion (Thompson et al. 1986; Perley 1999) that the radius of a facet (in radians) should be √θ 1 θfacet ≈ 3 where θ is the (FWHM) diameter of the synthesized beam (∼ λ/b for a baseline b). Larger facets can be used to reduce the computational burden at the expense of increased phase errors at the edges of the facets. (8) An alternate strategy, “targeted faceting,” exploits two aspects of the radio astronom- ical sky. First, the ma jority of the sidelobe confusion results from the strongest sources either inside or outside the primary beam. The variance in the flux density is given by the expectation value of S 2N (S ), which because N (S ) ∝ S x with x < 2 (equation 7) means that the variance is dominated by the strongest sources. Second, most of the the sky remains largely dark (empty of sources), even at these frequencies. Based on these facts, instead of tessellating the entire primary beam, which would take a large amount of time and com- puter power, one makes use of a priori knowledge of the field of view, e.g., from the NVSS or WENSS, to place small facets only where there are sources. We usually have a few hun- dred small facets per pointing. 18 Depending upon the sensitivity of one’s image and the desired image, this strategy can produce a useful computational savings. Recently, a new technique for wide-field imaging know as the ”w-pro jection” has been proposed. It pro jects the intrinsically 3-D data onto a 2-D plane with an appropriate Fernel-like convolution, and avoid the teidum involved with polyhedron imaging and targetted facetting. This technique is currently being tested at the VLA (Cornwell, Golap, & Bhatnagar 2006). 16 In general, snapshot observations are not viable with the 74 MHz VLA. The large primary beam and crowded fields means that a snapshot observation produces too few visibilities for an adequate representation of the field of view. The exception is for sources that are so strong, e.g., Cyg A, Vir A, Cas A, that their flux densities dominate the flux densities of other sources in the field of view. 17The aips task IMAGR and the special-purpose, low-frequency imaging task VLAFM implement polyhedral imaging. 18 Determining the location of the facets can be done by the aips task SETFC or the special-purpose, low-frequency imaging task VLAFM. – 32 – 5.3. Astrometry Even with the relatively large synthesized beam, compared to that attainable at higher frequencies, accurate astrometry at 74 MHz is desired. The astronomical motivations are varied but include spectral index studies, for which one wants to align images obtained at different frequencies (to the extent allowed by any frequency-dependent shifts within the source), and followup at other wavelengths, for which positional accuracies of 1′′ or better can be required. If self-calibration is used during phase calibration (§4.4), self-calibration will “freeze” the large-scale refraction but leave the image with an arbitrary absolute position (as in Figure 11). The NVSS or WENSS sources that appear in the image can then be utilized to re-register the astrometry to an accuracy of approximately 5′′ . If the ionosphere is treated as a phase-delay screen (§4.5), the astrometric source grid used in the procedure results in source positions comparable to that of the survey from which the astrometric grid was constructed (e.g., a few arcseconds for NVSS), provided that there are no other systematic effects. 6. Solar Effects, Ionospheric Weather, and Dynamic Scheduling A potential operational improvement in conducting 74 MHz observations (and low- frequency observations in general) would be to determine when ionospheric conditions are sufficiently quiescent to make useful observations. Currently, observations are scheduled on the telescope well in advance and proceed at the scheduled time. This procedure carries the risk that the ionosphere could be in a sufficiently disturbed state so as to preclude useful observations. A better approach would be to schedule the telescope in a “dynamic” fashion. In order to do this, one would have to conduct a test or identify a proxy observable that could establish the state of the ionosphere rapidly. If the ionosphere was disturbed so that 74 MHz observations would be unlikely to be successful, observations at a higher frequency could proceed with the 74 MHz observations being deferred until such time as the ionosphere is more amenable to correction via methods described here.19 Our experience suggests a number of ways in which such dynamic scheduling could be implemented. 19A similar observing strategy is in place for high frequency observations on the VLA (22 and 43 GHz) for which the dominant source of phase fluctuations is the troposphere. In this case, if the tropospheric conditions appear poor, the high frequency observations can be deferred with lower frequency observations being observed instead. – 33 – In general, the ionospheric phase effects that dominate calibration issues at low fre- quencies are a manifestation of highly unpredictable ionospheric weather conditions. The key consideration for low frequency observations is not the total thickness of the ionosphere, as measured by the TEC, but variations in TEC such as TIDs and smaller scale ionospheric structures (§4.5.1). Stable ionospheric conditions can occur at mid-day, when the TEC is highest. At the same time, highly disruptive ionospheric scintillations can occur during the middle of the night. Periods of particular ionospheric instability are [after] sunset and especially sunrise (Figure 14), but it is often easy to avoid observations during these times. Solar activity can affect 74 MHz observations both directly, due to radio emission from the sun that occur during the daytime, and indirectly, from ionospheric turbulence generated by manifestations of space weather linked to solar activity that can occur any time of the day or night. At 74 MHz the quiet Sun is a benign 2 kJy disk (compared to Cyg A at 17 kJy), and, at 30′ or larger in size, is significantly resolved out with the A and B configurations. However, nonthermal solar radio noise [storms] can have flux densities in excess of 1 MJy, and by entering through the far-out sidelibes of the primary beam render observations completely incapable of being calibrated. However they are usually short-lived, on the order of 30 minutes or less, and can be excised from the data in way similar to the treatment of narrow- band RFI. The second direct effect is due to the scattering effects of the solar wind. These lead to asymetric angular broadening and distortion of the brightness distribution of observed sources, and the effects are worst in the A and B configuration when the angular resolution is highest. Experience with both of these direct effects indicates that a useful rule of thumb is to allow at least 60 degree stand-off betweeen the target source and the sun. The indirect effects are more important since bad ionospheric weather conditions re- lated to a geomagnetic storm precipitated by solar activity can persist for days and render observations useless throughout. One well known manifestation are the massive solar ejec- tions of plasma and magnetic field known as Coronal Mass Ejections (CMEs). Energetic Earth-ward directed CMEs take 20 to 48 hours to arrive and, given a suitable magnetic field orientation, their impact can transfer massive amounts of energy into the upper atmosphere. The resulting geomagnetic storms can damage satellites, knock out power grids, and disrupt communications. Ionospheric disturbances of terrestrial origin, for example acoustic gravity waves can also generate ionospheric variations and poor low frequency observing conditions. However the most severe ionospheric disturbances are usually associated with solar activity. Various spacecraft and ground-based observatories monitor the Sun, and indices exist to – 34 – quantify the level of solar and geomagnetic activity. Hence proxies exist to predict both direct and indirect effects of solar and geomagnetic activity, and in principle these could be used to provide an ”ionospheric weather” forecast to guide low frequency observations. However no systematic effort has been conducted yet to assess the correlation between various indices and satisfactory ionospheric imaging conditions. In the mean time, the self-calibration solution, or short time scale stability of the phase on a few baselines from a short scan on a strong 74 MHz source (Cyg A, Vir A) is the most useful means of determining whether ionospheric weather conditions are suitable for observations. If they are deemed too poor to observe, the observations should be postponed, and conditions revisited on the timescale of several hours, or sooner if monitoring of strong sources can be efficiently built in to a dynamic scheduling system. Only limited attempts have been made to quantify the usefulness of this latter approach; while this procedure appears useful, it may not be able to distinguish excellent ionospheric imaging conditions from only fair conditions. 7. The Future of the 74 MHz System The 74 MHz system expanded from an initial trial system to an operational component of the VLA. Even so, its sensitivity is limited fundamentally by the VLA’s modest collecting area (∼ 104 m2 , which translates to ∼ 2 × 103 m2 given the ∼15% aperture efficiency at this frequency) and poor sensitivity. Even if the VLA telescopes could be made 50% efficient, and the bandwidths increased significantly (§3), it is not possible for the VLA (or the EVLA or the GMRT) to be more sensitive in plausible integration time (. 10 hrs) than roughly 1 mJy beam−1 at 74 MHz—100 times less sensitive than at 1400 MHz to normal-spectrum ob jects (i.e., those with spectral indices α < −0.7, Sν ∝ ν α ). The only way to achieve a sensitivity at these wavelengths comparable to that available at centimeter wavelengths is to build a system with much greater collecting area. The Long Wavelength Array (LWA20 , Kassim et al. 2006a,b; Taylor 2006; Kassim & Erickson 1998), Low-Frequency Array (LOFAR21 , Kassim et al. 2001, 2004; Falcke 2005) and Mileura Widefield Array (MWA22 , Morales et al. 2006; Bowman et al. 2007) seek to accomplish this at different levels, with different collecting areas, spatial resolutions and frequency ranges. The ob jective of these pro jects is to produce aperture synthesis instruments, operating at frequencies between 10 and 300 MHz, with collecting areas of up to ∼ 106 m2 at 20 MHz with 20 lwa.unm.edu 21www.lofar.org 22www.haystack.mit.edu/ast/arrays/mwa/ – 35 – maximum baselines of up to 500 km or longer. As such, they would have considerably more collecting area than the VLA (and more than any other previous low-frequency telescope as well) and angular resolutions comparable to that of the VLA at frequencies near 1 GHz. These pro jects must take into account the lessons learned from previous low frequency instruments. Here we emphasize some of these lessons, which became apparent during the deployment of the 74 MHz VLA system: • The system should be designed with careful attention to the signals it emits so that the instrument does not pollute itself with RFI. • The complex gain and noise temperature of each antenna should be monitored con- tinuously. Measurements of the round-trip phase through the system should also be available. • Channel bandwidths should be kept small (∆ν /ν ≪ 1) so as to avoid strong terrestrial transmitters as often occur at low frequencies and to enable wide-field imaging. • Wide-band receiving systems should be employed so that more than a single frequency can be observed. Compare the VLA to the CLRO in Figures 1 and 2. • Large, well-filled primary collectors (presumably composed of numerous individual dipole antennas phased together) should be used so as to produce a primary beam with high forward gain. • Ionospheric calibration should be considered during the design of the array. In particu- lar, adequate ionospheric calibration may require good aperture -plane coverage, rather than good u-v plane coverage usually sought in synthesis imaging. Also, the ability to make simultaneous measurements at widely spaced frequencies (for phase transfer) and pointing positions (for modelling the ionosphere over the array) should be avail- able. Rapid (¡ 10sec) switching may suffice over true simultaneous measurements, given sufficient sensitivity to track ionospheric changes on short time scales. 8. Conclusions We have described the 74 MHz system that is installed on the NRAO Very Large Array and the Pie Town antenna of the Very Long Baseline Array. All 29 antennas have been outfitted with 74 MHz receivers. With these low-frequency receivers and in the VLA’s A configuration, the VLA is the world’s highest resolution, highest dynamic range interferom- eter operating below 150 MHz. Working in conjunction with the PT antenna, the VLA-PT – 36 – interferometer, with baselines approaching 70 km, represents the longest baselines ever used for connected element synthesis imaging below 150 MHz. The calibration strategy for the 74 MHz VLA incorporates a number of features not common at higher frequencies. Chief among these is the importance of the isoplanatic patch size relative to the primary beam diameter. At 74 MHz, the isoplanatic patch size is deter- mined by the ionosphere and the size of the array, particularly in its larger configurations, and can be smaller than the primary beam. We have developed new field-based methods of calibrating 74 MHz VLA data that are not restricted to assuming that a single phase correction must apply to the entire field of view. For certain types of observations, e.g., a field dominated by a single strong source that itself is the ob ject of interest, normal, position- independent self-calibration (such as is used at centimeter-wavelengths) can be employed to obtain images with reasonable dynamic ranges. We have presented snapshot images of 3C sources of moderate strength as examples of routine, angle-invariant calibration and imag- ing, and a sub-sample of these sources with previously well determined low frequency spectra indicate that the 74 MHz flux scale at the Very Large Array is stable and reliable to at least 5 percent. The absolute flux density scale is tied to a model of Cygnus A with a flux density fixed to the Baars et al. (1977) value. Other aspects of calibration and imaging are not unique to 74 MHz but assume greater importance relative to higher frequencies. For instance, at 1400 MHz one can observe in a pseudo-continuum mode (i.e., low spectral resolution mode) in order to maintain a large field of view, to identify RFI, or both. At 74 MHz, the internal RFI is sufficiently bad that a pseudo-continuum mode (with somewhat higher spectral dynamic range than used typically at 1400 MHz) is essential. Similarly, at 1400 MHz there are typically other sources in the field of view, and it is possible for there to be strong sources outside the field of view that must be cleaned in order that their sidelobes do not reduce the dynamic range in the image. At 74 MHz, the presence of many sources in the field of view and of strong sources outside the field of view is guaranteed. Although we expect that the 74 MHz system will continue to be a productive, “facility- level” system of the VLA (and EVLA) for many more years, we also anticipate that the 74 MHz system will be superseded eventually by LOFAR and the LWA and possibly the low-frequency end of the Square Kilometer Array (SKA). Nonetheless, we believe that it offers many lessons for these future high resolution telescopes, as well as the apparatus to explore new science in its own right. We have benefited from discussions with many individuals. A partial listing includes M. Bietenholz, K. Blundell, C. Brogan, T. Clarke, J. Condon, K. Dyer, T. Ensslin, C. Lacey, – 37 – W. Tschager, and K. Weiler. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Uni- versities, Inc. ASC and WML were supported by National Research Council-NRL Research Associateships. Basic research in radio astronomy at the NRL is supported by 6.1 base funding. A. Snapshot Observations of Bright Sources The prototype 8-antenna 74 MHz system (Kassim et al. 1993) was capable of imaging only the dozen or so strongest sources in the sky whose flux densities were hundreds to thousands of Janskys. After the full deployment of the VLA 74 MHz system, a multi- configuration snapshot survey of sources from the 3C catalog was started. Observations were obtained in a succession of three VLA configurations in 1998, A configuration 7-8 March, B configuration 4-5 October, and C configuration 21 November and 4-5 December. Data in both circular polarizations was obtained simultaneously in 1 IF each at 74 and 330 MHz. The data were obtained in spectral line mode with 32 channels at 74 MHz and 64 channels at 330 MHz after online Hanning smoothing. The total available bandwidth was ∼1.5 and ∼3 MHz at 74 and 330 MHz, respectively. All sources were observed numerous times in cycling snapshot fashion to maximize the hour angle coverage. Our typical scan lengths were 5-10 minutes. Many of the sources were sufficiently small in angular extent that the A configuration run was sufficient to generate a good image. More extended sources required B and sometimes C array data, especially at 330 MHz, because of the higher intrinsic angular resolution. Table 7 summarizes the observations and the final image beams at 74 and 330 MHz. Notice that in most cases we convolved the final images with a gaussian, to produce a circular beam. The data reduction followed the prescriptions described in the previous sections of this paper. All of the images were produced from multiple snapshot observations using an angle- invariant calibration (self-calibration) strategy, which is sufficient for these strong sources because they dominate the self-calibration solution and sidelobe confusion can be ignored. In some cases, work is underway to produce even higher resolution, higher dynamic range images. Superior images are readily obtained with full synthesis observations as compared to the snapshot images presented here. Because of the calibration method used, the locations of the 330 and 74 MHz images are uncertain due to ionospheric wander. We used scaled subtractions (330 MHz − αave × 73 MHz) to test for shifts between the maps and adjusted the 74 MHz images to agree with – 38 – those at 330 MHz. Note that 330 MHz astrometry could be off by as much as 5′′ from the true radio reference frame. Alternatively, by using positions from the NVSS survey one can achieve positional accuracies of ∼1′′ at this frequency. Figure 20 shows 74 and 330 MHz images of a variety of moderately resolved 3C sources with flux densities on the scales of tens of Janskys or higher. Images such as these can now be made routinely with snapshot observations of tens of minutes or less. In most cases these are the first sub-arcminute images of these sources below 100 MHz. Table 8 reports the peak brightness and flux density of the source, but, for the resolved sources for which we present images, these values may be lower limits, as some of the flux may have been resolved out. We discuss the resolved sources briefly but make no quantitative analyzes, as our discussion here is intended only to be illustrative. 3C 10 Figure 20a shows the images of Tycho’s supernova remnant at 74 and 330 MHz. We can see in these images a limg-brightened spherical shell with a diameter of ∼8′ , and enhanced emission toward the NE half of the shell. This structure is similar to that observed at 330 MHz and 1.4 GHz by Katz-Stone et al. (2000), and at 610 MHz by Duim & Strom (1975). 3C 33 Figure 20b presents the 330 and 74 MHz images of this radio galaxy. The 74 MHz image shows the lobes and fainter emission between these two structures. The 330 MHz image resolves some of the details of this source, separating each lobe into 2 compo- nents, as well as showing diffuse emission extending perpendicular to the jet axis in regions between the lobes and the nucleus. The structures seen in our images are sim- ilar to the ones seen at 1.5 GHz by Leahy, & Perley (1991). A low resolution 160 MHz image of this source (Slee 1977) was able to resolve it only into two components. 3C 84 Figure 20c shows the 74 MHz image of the central regions of the Perseus cluster, which contains a number of strong radio sources including 3C 84. An enlarged plot of the 330 MHz Figure 20d image of this radio galaxy shows a strong core, two lobes in the N-S direction and some extended emission beyond the lobes. These structures are similar to the ones detected by Pedlar, A. et al. (1990) at 1.4 GHz, 330 MHz and 151 MHz. The 74 MHz image does not resolve the nucleus and lobes, but shows some diffuse emission extending to the NW and SW, corresponding to previous outbursts (see Fabian et al. (2002) for a more detailed discussion of this image and the relation between these structures and X-ray holes). 3C 98 Figure 20e The 74 MHz image shows a double lobe structure with diffuse emission in between. The 330 MHz image shows a similar structure, although with higher resolution, allowing the detection of the hotspots and part of the jet. These structures – 39 – were previously imaged at higher frequencies (1.4 and 4.8 GHz) by Leahy et al. (1997) and Young et al. (2005). A 160 MHz image of this galaxy was presented by Slee (1977). 3C 129 Figure 20f shows 3C 129 at 74 MHz. A study of the 330 MHz image (Lane et al. 2002) have confirmed the existence of a steep-spectrum “crosspiece” at the head of 3C 129, along the NE-SW direction, perpendicular to its main tail. This structure was previously detected at 600 MHz (Jagers & de Grijp 1983) and have recently been observed by Lal & Rao (2004) with the GMRT at 240 and 610 MHz. We see no indication of this structure at 74 MHz, but our resolution is considerably lower, so it is probably blended with the head of 3C 129. 3C 144 Figure 20g shows 3C 144 (the Crab Nebula, Tau A) at 74 MHz and 330 MHz. This source can be used as an amplitude calibrator in more compact configurations (C and D). The 74 MHz compact source in the center of the nebula is the Crab pulsar (PSR B0531+21). A higher sensitivity 330 MHz image of this source is presented by Frail et al. (1995), while Bietenholz et al. (1997) present a study of the 74/330 MHz radio spectral index and find evidence for intrinsic thermal absorption. 3C 218 (Hydra A) Figure 20h present the 330 and 74 MHz images of Hydra A, where we can see that this radio galaxy has a complex structure, consisting of several outbursts. Lane et al. (2004a) present a detailed study of this source, the spectral indices of the different components and their correlation of the X-ray emission from the cluster of galaxy where it resides. 3C 219 The 74 MHz emission from this radio galaxy (Figure 20i ) shows two lobes and diffuse emission between them. This emission is resolved into better detail by the 330 MHz image, where we can see the N and S hotspots, the jet and some diffuse emission. These structures are similar to the ones seen at 1.4 GHz by Clarke et al. (1992). 3C 274 (Vir A) Figure 20j shows 3C 274 (Vir A), one of the sources that can be used as a primary bandpass and flux density calibrator, in addition to or instead of Cyg A (particularly if Cyg A is not above the horizon). Both 330 MHz and 74 MHz images show a complex structure, indicative of the interaction between the radio plasma and the intra cluster medium. A more detailed discussion about the 330 MHz image of this source is presented by Owen et al. (2000). 3C 327 The 74 MHz image of this galaxy (Figure 20k ) shows a double structure along the E-W direction. The E component is elongated, indicating the presence of multiple components. A low resolution 160 MHz image from Slee (1977) showed only two blobs – 40 – along the E-W direction. The structure seen at 74 MHz are confirmed by the 330 MHz image, where we can clearly see the hotspots and some diffuse emission, similar to the structure observed at 8.4 GHz by Leahy et al. (1997). 3C 353 The 74 MHz image of this radio galaxy (Figure 20l ) is broken into 3 components aligned along the E-W direction. Previous low frequency (160 MHz) images by Slee (1977) were not able to detect multiple components, detecting only an extended source in the E-W direction. The 330 MHz image shows the hotspots and diffuse emission in the lobes, similar to the structure detected by Baum et al. (1988) at 5 GHz. 3C 390.3 In Figure 20m we can see that the 74 MHz image of this radio galaxy is resolved into 2 lobes and diffuse emission associated with them. This structure is similar to the one detected at 610 MHz by radio Jagers (1987), using WSRT observations. The 330 MHz image shows the hotspots and diffuse emission in better detail, as well as the nucleus. The structures seen at this frequency are similar to the ones detected with 1.4 GHz VLA observations (Leahy & Perley 1995). 3C 392 Figure 20n presents the images of this supernova remnant, which has similar struc- ture at 74 and 330 MHz. The 330 MHz image shows details similar to the ones seen at 1.4 GHz (Jones, Smith & Angellini 1993; Giacani et al. 1997). Higher resolution, full synthesis images at both frequencies of this SNR, also known as W44, are presented in Castelletti et al. (2007). 3C 405 (Cyg A) Figure 20o shows our primary bandpass and flux density calibrator, 3C 405 or Cyg A. This image is dynamic range limited. 3C405 is almost unresolved at 74 MHz, however, higher resolution (VLA + Pie Town) images at both 74 and 330 MHz are presented by Lazio et al. (2006). The VLA 330 MHz image is able to resolve the emission into hotspots and associated diffuse emission, similar to the struc- ture detected by Perley & Erickson (1984) at 1.4 GHz, although without the detection of the nucleus and the jets. 3C 445 Figure 20p shows that 74 MHz image of this radio galaxy is composed of two bright lobes and some faint extended emission around the nucleus. A lower resolu- tion 160 MHz image of this galaxy (Slee 1977) shows only three blobs. The VLA 330 MHz image resolves the lobes into hotspots and some associated diffuse emission, similar to that observed at higher frequencies (Kronberg, Wielebinski & Graham 1986; Leahy et al. 1997). 3C 452 The 74 MHz image of this radio galaxy is presented in Figure 20q, which shows a double lobed structure, similar to the one detected at 610 MHz by Jagers (1987), – 41 – using WSRT. The 330 MHz image resolves the structure of this galaxy into better detail, hotspots and diffuse emission along the jet, similar to the structure detected at 1.4 GHz by Dennett-Thorpe, et al. (1999). 3C 461 Figure 20r shows the 74 and 330 MHz of Cas A, another source that can be used as a primary bandpass and flux density calibrator in more compact configurations (C and D). The 330 and 74 MHz images look similar, although of lower resolution compared to the 1.4 GHz images presented in Anderson et al. (1991). Kassim et al. (1995) found evidence of internal thermal absorption using the prototype 8-antenna system, that was subsequently confirmed by Delaney et al. (2004) using the full 74 MHz VLA + PT Link (Figure 21). These observations were obtained with a maximum baseline is 72 km, corresponding to an angular resolution of 8.5′′ . Alfv`en, H. & Herlofson, N. 1950, Phys. Rev., 78, 616 REFERENCES Anderson, M., Rudnick, L., Leppik, P., Perley, R., & Braun, R. 1991, ApJ, 373, 146 Baars, J. W. M., Genzel, R., Pauliny-Toth, I. I. K., & Witzel, A. 1977, A&A, 61, 99 Baum, S. A., Heckman, T. M., Bridle, A., van Breugel, W. J. M., & Miley, G. K. 1988, ApJS, 68, 643 Bietenholz, M. F., Kassim, N., Frail, D. A., Perley, R. A., Erickson, W. C., & Ha jian, A. R. 1997, ApJ, 490, 291 Bowman, J. D. et al. 2007, AJ, in press Braude, S. Ia., Men, A. V., & Sodin, L. G. 1978, Antenny, 26, 3 Brentjens, M. A., 2005, Astron. Nachr., 326, 609-609 Bridle, A. H. 1999, in Synthesis Imaging in Radio Astronomy, eds. R. A. Perley, F. R. Schwab, & A. H. Bridle (ASP: San Francisco) p. 443 Bridle, A. H. & Purton, C. R. 1968, AJ, 73, 717 Bock, D. C.-J., Large, M. I., & Sadler, E. M. 1999, AJ, 117, 1578 Castelletti, G., Dubner, G., Brogan, C., Kassim, N. E. A&A, in press (astro-ph/0702746) – 42 – Clarke, D. A., Bridle, A. H., Burns, J. O., Perley, R. A., & Norman, M. L. 1992, ApJ, 385, 173 Cohen, A.S., Lane, W.M., Kassim, N.E., Lazio, T.J.W., Cotton, W.D., Condon, J.J., Perley, R.A., Erickson, W.C. 2006, in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science”, ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA. Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p.299. Cohen, A. S., Rottgering, H. J. A., Kassim, N. E., et al. 2003, ApJ, 591, 640 Cohen, A. S., et al. 2007, in preparation Condon, J. J., Cotton, W. D., Greisen, E. W., Yin, Q. F., Perley, R. A., Taylor, G. B., & Broderick, J. J. 1998, AJ, 115, 1693 Cornwell, T.J., Golap, K., Bhatnagar, S. in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science”, ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA.Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p.350. Cornwell, T. J. & Fomalont, E. B. 1999, in Synthesis Imaging in Radio Astronomy II, eds. G. B. Taylor, C. L. Carilli, & R. A. Perley (ASP:San Francisco) p. 187 Cornwell, T. J. & Perley, R. A. 1992, A&A, 261, 353 Cotton, W.D. in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science” ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA. Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p.337 Cotton, W. D. & Condon, J. J. 2002, Proc. URSI General Assembly, J3.0.2 Cotton, W. D., Condon, J. J., Perley, R. A., Kassim, N., Lazio, J., Cohen, A., Lane, W., & Erickson, W. C. 2004, Proc. SPIE, 5489, 180 Delaney, T., Rudnick, L., Jones, T., Fesen, R., Hwang, U., Petre, R., & Morse, J. 2004, in X- Ray and Radio Connections Conference Proceedings, eds. L.O. Sjouwerman and K.K. Dyer. Published electronically by NRAO, http://www.aoc.nrao.edu/events/xraydio . Held 3/6 February 2004 in Santa Fe, New Mexico, USA (E4.05) Dennett-Thorpe, J., Bridle, A. H., Laing, R. A., & Scheuer, P. A. G. 1999, MNRAS, 304, 271 – 43 – Duim, R. M., & Strom, R. G. 1975, A&A, 39, 33 Erickson, W. C. 1984, J. Astrophys. Astron., 5, 55 Erickson, W. C., Perley, R. A., Flatters, C., & Kassim, N. E. 2001, A&A, 366, 1071 Erickson, W. C., Mahoney, M. J., & Erb, K. 1982, ApJS, 50, 403 Erickson, W.C. in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science”, ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA. Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p.317 Fabian, A. C., Celotti, A., Blundell, K. M., Kassim, N. E., & Perley, R. A. 2002, MNRAS, 331, 369 Falcke, H. 2005, Astron. Nachr., 326, 612 Frail, D. A., Kassim, N. E., Cornwell, T. J., & Goss, W. M. 1995, ApJ, 454, L129 Giacani, E. B., Dubner, G.. M., Kassim, N. E., Frail, D. A., Goss, W. M., Winkler, P. F., & Williams, B. F. 1997, AJ, 113, 1379 Gizani, Nectaria A. B., Cohen, A., Kassim, N. E., 2005 MNRAS, 358, 1061 Golap, K., Cornwell, T.J., Perley, R.A., Bhatnagar, S. in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science”, ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA. Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p.366. Hewish, A., Bell, S. J., Pilkington, J. D. H., Scott, P. F., & Collins, R. A. 1968, Nature, 217, 709 Jansky, K. G. 1935, Proc. I. R. E., 23, 1158 Jacobson, A. R. & Erickson, W. C. 1992a, Planetary and Space Sci., 40, 447 Jacobson, A. R. & Erickson, W. C. 1992b, A&A, 257, 401 Jagers, W. J. 1987, A&AS, 67, 395 Jagers, W. J., & de Grijp, M. H. K. 1983, A&A, 127, 235 Jones, L. R., Smith, A., & Angellini, L. 1993, MNRAS, 265, 631 – 44 – Kassim, N. E., & Erickson, W. C. 1998, Proc. SPIE, 3357, 740 Kassim, N. E., et al. 2006, Long Wavelength Astrophysics, 26th meeting of the IAU, Joint Discussion 12, 21 August 2006, Prague, Czech Republic, JD12, #56, 12 Kassim, N. E., Lazio, T. J. W., Erickson, W. C., et al. 2000, in Radio Telescopes, Proc. SPIE, ed. H. R. Butcher, vol. 4015, p. 328 Kassim, N. E., Perley, R. A., Dwarakanath, K. S., Erickson, W. C. 1995, ApJ, 455, L59 Kassim, N. E., Perley, R. A., Erickson, W. C., & Dwarakanath, K. S. 1993, AJ, 106, 2218 Kassim, N. E., Lazio, T. J. W., Ray, P. S., Crane, P. C., Hicks, B. C., Stewart, K. P., Cohen, A. S., Lane, W. M., 2004, Planetary and Space Science, Volume 52, Issue 15, p. 1343-1349 Kassim, N. E. 1988, ApJS, 68, 715 Kassim, N. E., Polisensky, E. J., Clarke, T. E., Hicks, B. C., Crane, P. C., Stewart, K. P., Ray, P. S., Weiler, K. W., Rickard, L. J., Lazio, T. J. W., Lane, W. M., Cohen, A. S., Nord, M. E., Erickson, W. C., Perley, R. A. in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science”, ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA. Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p.392 Katz-Stone, D. M., Kassim, N. E., Lazio, T. J. W., O’Donnell, R. 2000, ApJ, 529, 453 Kiepenheuer, K. O. 1950, Phys. Rev., 79, 738 Kronberg, P. P., Wielebinski, R., & Graham, D. A. 1986, A&A, 169, 63 Krymkin, V. V. & Sidorchuk, M. A. 1994, Ap&SS, 213, 1 Kuhr, H., Witzel, A., Pauliny-Toth, I. I. K., & Nauber, U. 1981, A&AS, 45, 367 Lane, W. M., Clarke, T. E., Taylor, G. B., Perley, R. A., Kassim, N. E. 2004, AJ, 127, 48 Lane, W. M., Kassim, N. E., Ensslin, T. A., Harris, D. E., & Perley, R. A. 2002, AJ, 123, 2985 Lane, W., Cohen, A., Cotton, W. D., Condon, J. J., Perley, R. A., Lazio, J., Kassim, N., Erickson, W. C. in ”Ground-based Telescopes”. Edited by Oschmann, Jacobus M., Jr. Proceedings of the SPIE, Volume 5489, pp. 354-361 (2004) – 45 – Lane, W. M., Cohen, A. S., Kassim, N. E., Lazio, T. J. W., Perley, R. A., Cotton, W. D., Greisen, E. W., Radio Science, 40, RS5S05, 2005. (See also Golap et al. 2006.) Lane, W.M., Cohen, A.S., Kassim, N.E., Lazio, T.J.W., in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science”, ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA. Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p.203. Lal, D.V., & Rao, A. P. 2004, A&A, 420, 491 Lazio, T. J. W., Cohen, A. S., Kassim, N. E., Perley, R. A., Erickson, W. C., Carilli, C. L., & Crane, P. C. 2006, ApJ, 642, L33 Leahy, J. P., Black,A. R. S., Dennett-Thorpe, J., Hardcastle, M. J., Komissarov, S., Perley, R. A., Riley, J. M., & Scheuer, P. A. G. 1997, MNRAS, 291, 20 Leahy, J. P., Muxlow, T. W. B., & Stephens, P. W. 1989, MNRAS, 239, 401 Leahy, J. P. & Perley, R. A. 1991, AJ, 102, 537 Leahy, J. P., & Perley, R. A. 1995, MNRAS, 277, 1097 Mills, B. Y. 1952, Aust. J. Sci. Res., A5, 456 McCready, L. L,. Pawsey, J. L., & Payne-Scott, R. 1947, Proc. Royal Soc. A, 190, 357 Morales, M.F., Lonsdale, C.J., Cappallo, R.J., Hewitt, J.N., Doeleman, S., in ”From Clark Lake to the Long Wavelength Array: Bill Erickson’s Radio Science”, ASP Conference Series, Vol. 345, Proceedings of the Conference held 8-11 September, 2004 in Santa Fe, New Mexico, USA. Edited by N. Kassim, M. Perez, M. Junor, and P. Henning, p. 452. Nijboer, R. J., Noordam, J. E., Yatawatta, S. B. in ”Astronomical Data Analysis Software and Systems XV” ASP Conference Series, Vol. 351, Proceedings of the Conference Held 2-5 October 2005 in San Lorenzo de El Escorial, Spain. Edited by Carlos Gabriel, Christophe Arviset, Daniel Ponz, and Enrique Solano. San Francisco: Astronomical Society of the Pacific, 2006., p.291 Noordam, J.E. 2004, in ”Ground-based Telescopes”. Edited by Oschmann, Jacobus M., Jr. Proceedings of the SPIE, Volume 5489, pp. 817-825 Pedlar, A., Ghataure, H. S., Davies, R. D., Harrison, B. A., Perley, R., Crane, P. C., & Unger, S. W. 1990, MNRAS, 246, 477 – 46 – Napier, P. J., Thompson, A. R., & Ekers, R. D. 1983, Proc. IEEE, 71, 1295 Owen, F. N., Eilek, J. A., & Kassim, N. E. 2000, ApJ, 543, 611 Pawsey, J. L., Payne-Scott, R., & McCready, L. L. 1946, Nature, 157, 158 Perley, R. A. 2002, Proc. URSI General Assembly, J4.0.2 Perley, R. A. 1999, in Synthesis Imaging in Radio Astronomy II, eds. G. B. Taylor, C. L. Car- illi, & R. A. Perley (San Francisco: ASP) p. 383 Perley, R. A. & Erickson, W. C. 1984, VLA Scientific Memorandum #146 Reber, G. 1940, ApJ, 91, 621 Rees, N. 1990, MNRAS, 244, 233 Rengelink, R. B., Tang, Y., de Bruyn, A. G., Miley, G. K., Bremer, M. N., Rottgering, H. J. A., & Bremer, M. A. R. 1997, A&AS, 124, 259 Roger, R. S., Costain, C. H., & Stewart, D. I. 1986, A&AS, 65, 485 Ryle, M. 1952, Proc. Royal Soc. A, 211, 351 Ryle, M., Smith, F. G., & Elsemore, B. 1950, MNRAS, 110, 508 Ryle, M. & Vonberg, D. D. 1946, Nature, 158, 339 Shepherd, D. S., Claussen, M. J., & Kurtz, S. E. 2001, Science, 292, 1513 Shkolvsky, I. S. 1952, AZh, 29, 418 Slee, O. B. 1977, Au. J. Phys. Astron., 43, 1 Slee, O. B. & Higgins, C. S. 1975, Au. J. Phys. Astron., 36, 1 Slee, O. B. & Higgins, C. S. 1973, Au. J. Phys. Astron., 27, 1 Subramanian, K. R., Gowda, C. N., Hameed, A. T. A., & Sastry, C. V. 1986, Astron. Soc. India. Bulletin, 14, 236 Swarup, G. 1990, Indian Journal of Radio and Space Physics, 19, 493 Taylor, G. B. 2006, Long Wavelength Astrophysics, 26th meeting of the IAU, Joint Discussion 12, 21 August 2006, Prague, Czech Republic, JD12, #17, 12, – 47 – Thomasson, P. 1986, QJRAS, 27, 413 Thompson, A. R., Moran, J. M., & Swenson, G. W., Jr. 1986, Interferometry and Synthesis in Radio Astronomy (Wiley: New York) Walker, R. C. 1995, in Very Long Baseline Interferometry and the VLBA, eds. J. A. Zensus, P. J. Diamond, & P. J. Napier (ASP:San Francisco) p. 247 Yates, K. W. 1968, Au. J. Phys., 21, 167 Young, A., Rudnick, L., Katz, D., DeLaney, T., Kassim, N. E., & Makishima, K. 2005, ApJ, 626, 748 This preprint was prepared with the AAS LATEX macros v5.2. – 48 – DRAO DRAO UTR-2 Gauribidanur CLRO 8C Culgoora VLA Mauritius Culgoora GMRT Merlin 10 100 Frequency (MHz) 1000 100 10 1 Fig. 1.— Angular resolution (arcseconds) as a function of frequency (MHz) for past and present imaging instruments in the 10 to 200 MHz range. The various telescopes include the UTR-2 (Krymkin & Sidorchuk 1994), Gauribidanur (Subramanian et al. 1986), 8C (Rees 1990), Clark Lake Radio Observatory (CLRO, Kassim 1988), Culgoora (Slee & Higgins 1973, 1975; Slee 1977), Dominion Radio Astrophysical Observatory (DRAO, Bridle & Purton 1968; Roger et al. 1986), MERLIN (Thomasson 1986; Leahy, Muxlow & Stephens 1989) and the Mauritius radio telescope. The 74 MHz VLA is respresented by the filled triangle. – 49 – 1000 100 10 1 0.1 DRAO DRAO Gauribidanur UTR-2 CLRO 8C Culgoora Culgoora VLA Mauritius Merlin GMRT 10 100 Frequency (MHz) Fig. 2.— Sensitivity (mJy) to a point source as a function of frequency for the same instru- ments shown in Figure 1. The sensitivities are estimates of the minimum detectable flux density provided by past and present telescopes. The sensitivity of most of the telescopes shown here were or are confusion limited; for the VLA and the GMRT, an integration time of 8 hr was assumed in calculating their sensitivities. Fig. 3.— A picture of a 74 MHz dipole mounted on a VLA antenna. In the center of the picture is the subreflector, supported by the quadrupod legs. The 74 MHz (crossed) dipoles are in the lower center of the picture. The cable that carries the signals from the dipoles to the receivers drops from the intersection of the dipoles to the bottom of the antenna’s surface. Also visible just below the subreflector are the 330 MHz dipoles. Fig. 4.— The block diagram of the 74 MHz receiver on the Pie Town VLBA antenna. The receivers on the VLA antennas are similar, with the main difference being that the VLA receivers are not hetrodyne receivers. Rather the 74 and 330 MHz signals are transferred directly to the intermediate frequency (IF) transmission system. – 50 – ) s e e r g e D ( e s a h P a n n e t n A 60 40 20 0 -20 10.4 ) s e e r g e D ( e s a h P a n n e t n A 500 400 300 200 100 0 -100 -200 10.4 VLA Antenna 7 at E16 10.6 Time (MST, Hours) 10.8 VLBA Antenna at PieTown 10.6 Time (MST, Hours) 10.8 Fig. 5.— First fringes using the PT Link at 74 MHz. The source observed was 3C 123. Top Phase as a function of time for VLA antenna #7, located relatively close to the center of the array. Bottom Phase as a function of time for the PT antenna. In both cases, the phases are measured relative to an antenna near the center of the array. The scales on the ordinates differ. Fig. 6.— The primary beam power pattern at 74 MHz from one of the antennas. Other antennas have similar power patterns. The left panel shows the left circular polarization, and the right panel shows the right circular polarization. The axes are the sines of the offset angle. The contours show the decrease, in dB, from the peak of the pattern (drawn at 3, 6, 10, 15, 20, 25, and 30 dB levels). The sharp edge on the right hand side of the plots results from the motion limits on pointing for the VLA antennas. – 51 – Fig. 7.— The rms image noise level as a function of receiver bandwidth. The dots show the measured values from a 90 min. integration calibrated and imaged in the fashion described in §§4 and 5. The solid line shows the behavior expected, ∆ν −1/2 , if the system performance is limited by thermal noise. This behaviour is typical of observations obtained in the A and B configurations. – 52 – Fig. 8.— The rms image noise level as a function of integration time. Multiple points at the same integration time indicate integration times for which multiple, independent images were constructed from the approximately 1-hr total integration time. The solid line shows the behavior expected, t−1/2 , if the system performance is limited by thermal noise. This behaviour is typical of observations obtained in the A and B configurations. – 53 – Fig. 9.— A demonstration of the excision of 74 MHz RFI. Left: The visibility amplitudes on a single baseline are displayed in a (linear) gray scale format with frequency on the abscissa and time on the ordinate. The time axis is not linear, and the thick horizontal black stripes indicate times when other sources were being observed. The bright vertical stripes represent the 100 kHz “comb”. The strength of the comb varies from baseline to baseline and as a function of time within individual baselines, depending upon the orientation of the antennas and how strongly they couple. For this illustration, we have chosen a particularly severe example; there were other baselines in the same data set for which the comb was barely visible. Right: The same data after RFI have been flagged with FLGIT. This panel shows a much more homogeneous brightness distribution than the one to the left, indicating that most of the RFI (strongest interference) was removed. – 54 – ) 0 5 9 1 B ( N O I T A N I L C E D 25 35 30 25 20 15 10 05 00 24 55 13 30 00 29 30 00 28 30 00 RIGHT ASCENSION (B1950) 27 30 00 Fig. 10.— An example of the distortions introduced by imaging a region larger than the isoplanatic patch. This image is a subimage of a larger image. The sources shown are roughly 7◦ from the phase center of the larger image; the two brightest sources should be point-like or nearly so while the third, fainter source is extended, with a nearly north-south orientation. The image was produced from combined B- and C-configuration observations; the beam is 95′′ × 83′′ and is shown in the lower left. The rms noise level in the image is approximately 30 mJy beam−1 , and the contour levels are 30 mJy beam−1 × − 4, −2, 2, 4, 6, 10, 20, 40, 60, 100, and 150. The bandwidth and time averaging used in producing this image are sufficiently small that both contribute a negligible amount (< 1 beamwidth) of smearing. – 55 – Fig. 11.— An illustration of the position offsets arising from phase distortions caused by the largest spatial structures (& 1000 km) in the ionosphere. The contours show the 74 MHz image; the gray scale is an overlay of the NVSS in this region. The NVSS is at a sufficiently high frequency that the ionospheric refraction is much less than a beamwidth. Note that not all NVSS sources have a 74 MHz counterpart. Fig. 12.— The temporal variation in the refractive shift caused by the largest spatial scales in the ionosphere. Shown is the offset, in both right ascension and declination, of Cyg A from its known position as a function of time, with a 1 min. sampling interval. These observations were taken in the B configuration. Fig. 13.— An illustration of the phase distortions caused by ionospheric mesoscale structure, in this case a traveling ionospheric disturbance (TID), with the typical scale of order hundreds of kilometers. Shown are the phases, measured relative to an antenna near the center of the array, as a function of time for three antennas along the west arm of the array. – 56 – Scintillation Refractive wedge At dawn ‘Midnight wedge’ Quiesence TIDs Fig. 14.— a) The phase of three antennas relative to a central antenna during an approxi- mately 8 hr observations of Vir A illustrating many of the ionospheric phenomena typically observed at the VLA. All three antennas are located approximately 8 km from the reference antenna, but represent different azimuths. For the TID the observed parameters were a pe- riod of 750i s, phase slope of 50 deg km−1 , and a time lag of ∼50 seconds over 20 km allowing Perley (2002) to derive a TID wavelength of 750 km and velocity of 200 m s−1 ; b) Same as in panel a except for two antennas at different distances along the same azimuth, indicating that to first order the phase effects of all the phenomena are proportional to baseline length; c) The instantaneous amplitude or apparent defocusing of Vir A over the same time scale as in panel a, (Hour Angle -4 corresponds to IAT Time∼8 hrs), sunrise occurred at +3h ; d) The refraction (or apparent postion wander in both RA and Dec) of Virgo A over the same time scale as panel a; e,f ) The apparent differential refraction in RA (e) and declination (f ) as measured towards 5 ob jects located within 6 degrees of Vir A and of sufficient strength to be detected and tracked over the course of the observations. The time scale is the same as in panel c. – 57 – – 58 – ) 0 0 0 1 o t d e z i l a m r o N ( y t i s n e D x u l F k a e P 1000 800 600 400 200 -4 -2 2 0 Hour Angle (Hr) 4 – 59 – – 60 – ) s d n o c e s c r a ( t e s f f O n o i s n e c s A t h g i R 300 200 100 0 -100 -200 -300 Differential Wander in Virgo A Field Right Ascension Shifts for Five Background Objects D = 5.4 deg, PA = 125 D = 6.5 deg, PA = 59 D = 5.1 deg, PA = 39 D = 3.9 deg, PA = -139 D = 4.3 deg, PA = -67 -4 -2 0 Hour Angle (Hr) 2 4 – 61 – ) s d n o c e s c r a ( t e s f f O n o i t a n i l c e D 300 200 100 0 -100 -200 -300 Differential Wander in Virgo A Field Declination Shifts for Five Background Objects D = 5.4 deg, PA = 125 D = 6.5 deg, PA = 59 D = 5.1 deg, PA = 39 D = 3.9 deg, PA = -139 D = 4.3 deg, PA = -67 -4 -2 0 Hour Angle (Hr) 2 4 – 62 – 00 -02 -04 -06 -08 ) 0 0 0 2 J ( N O I T A N I L C E D -10 02 50 40 30 20 RIGHT ASCENSION (J2000) 10 00 Fig. 15.— An example of the bias introduced by self-calibration. Symbols show the location of sources and are proportional to their flux densities. This field contains the (strong) source 3C 63 in the upper right. A clear non-uniform distribution of other sources across the field is evident, in addition to the decrease in number density expected from the primary beam attenuation. The jagged edges in the image indicate the boundaries of facets (§5); regions beyond the edges of the image have not been imaged. – 63 – Fig. 16.— Differential, ionospheric-induced source wander within a field of view. The ex- pected locations of various moderately strong sources within a single field of view is shown. The direction of each vector indicates direction of the position shift in five 5-min. intervals; the length of each vector is 100 times the actual displacement. Because the magnitude and direction of the wander is not the same for all sources, averaging over time results in sources being smeared out and apparently disappearing from view (viz. Figure 15). – 64 – -10 S E E R G E D -8 -6 -4 -2 0 2 4 6 8 10 10 C E S C R A C E S C R A C E S C R A 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 200 150 100 50 0 -50 -100 -150 -200 ARC SEC 200 150 100 50 0 -50 -100 -150 -200 ARC SEC 200 150 100 50 0 -50 -100 -150 -200 ARC SEC 5 0 DEGREES -5 -10 Fig. 17.— Left : A mosaic of NVSS sources assembled from a 1 min. snapshot of a full-field image at 74 MHz. The typical distance of each source from the phase center is 3◦ . The cross in each panel marks the nominal location of the NVSS source. Offsets from this nominal position are due to ionospheric refraction. Right : The Zernike model for φion,lo over the VLA for the same time as the mosaic was constructed. See equation (6). Shown is the phase delay screen above one VLA antenna; because of the “small-array” approximation used, the phase delay screen over all other antennas is essentially the same. At a typical 350 km altitude, the phase delay screen shows structures on scales of roughly 100 km. Solid contours represent a phase advance, relative to a nominal phase, while dashed contours represent a phase delay. The most negative contour (lower right) represents a phase delay of −30◦ , and the most positive contour (upper left) represents a phase advance of +30◦ . Fig. 18.— The same field as in Figure 15, but calibrated by treating the ionosphere as a phase-delay screen. The distribution of sources across the field is far more uniform than before, though the primary beam attenuation near the edges of the image is still apparent. Abscissa: Right ascension; Ordinate: Declination. The jagged edges in the image indicate the boundaries of facets (§5); regions beyond the edges of the image have not been imaged. The black circle outside the field of view in the upper right is the location of an outlier field; this field contains a strong source outside the field of view that was also imaged. – 65 – Fig. 19.— The rms jitter in apparent separations of pairs of sources with different separations made from a sequence of one minute snapshots taken over a period of several hours on a day with a moderately disturbed ionosphere. Top The rms jitter with no correction to the position difference. Bottom The rms jitter after applying an ionospheric model described by a 5-term Zernike phase screen determined from that snapshot. The large rms values at angular separations greater than approximately 15◦ result from Cyg A which was not corrected as part of this Zernike modeling. – 66 – Fig. 20.— Various (resolved) 3C sources at 74 MHz (left panel) and 330 MHz (right panel). For each source, we quote the rms noise level and the beam diameters in Table 7. All beams are circular unless otherwise noted. The contour levels are given in terms of the rms noise level in the image i(−3, 3, 6, 12, 24, 48, . . . times that rms noise level) and the beam is shown in the lower left. (a ) 3C 10; (b ) 3C 33; (c ) The central strong sources of the Perseus cluster at 74 MHz. The source in the lower left is 3C 84, the compact source near the center is 3C 83.1A, and the extended radio galaxy in the upper right is NGC 1265; (d ) 3C 84; (e ) 3C 98; (f ) 3C 129 (western source) and 3C 129.1 (eastern source); (g ) 3C 144; (h ) 3C 218; (i ) 3C 219; (j ) 3C 274 or Vir A; (k ) 3C 327; (l ) 3C 353; (m ) 3C 390.3; (n ) 3C 392; (o ) 3C 405 or Cyg A; (p ) 3C 445; (q ) 3C 452; (r ) 3C 461. Fig. 21.— VLA + PT Link full synthesis 74 MHz image of Cas A, with a resolution of ∼8′′ . The bottom bar shows the color transfer function in Jy. Image courtesy T. Delaney Delaney et al. (2004) – 67 – Table 1. Performance Characteristics of the VLA 74 MHz System Center Frequency Bandwidth Primary Beamwidth (FWHM) System Temperature (minimum) Aperture Efficiency Point Source Sensitivity (A or B configs., 8 hr) Resolution (A configuration) (VLA + PT) 73.8 MHz 1.6 MHz 11.◦7 1500 K ≈ 15% 25−50 mJy 25′′ 12′′ Note. — See also Figures 6, 7, and 8. – 68 – Table 2. Comparison of VLA and Kuhr et al. Flux Density Estimates Name Kuhr et al. (Jy) VLA (Jy) Ratio Remarks 3C 48 3C 98 3C 123 3C 147 69.1 98.9 387.6 67.1 3C 219 96.6 3C 274 3C 327 3C 353 3C 390.3 2281.3 93.1 437.0 107.2 3C 445 62.6 67.6 ± 0.4 98.0 ± 1.1 414.7 ± 1.3 55.3 ± 0.8 94.8 ± 0.5 2084.6 ± 1.3 118.4 ± 0.8 443.8 ± 0.8 97.1 ± 0.4 49.2 ± 0.3 0.978 Kuhr et al. spectrum fit below observations 0.991 1.070 Both Kuhr et al. spectrum and VLA data look good 0.825 Kuhr et al. estimate could be high, only one datum below 178 MHz and spectrum fit well below it 0.981 Kuhr et al. spectrum looks good 0.914 Both Kuhr et al. spectrum and VLA data look good 1.272 Kuhr et al. spectrum could be low, falls below 38 MHz datum but fits 80 MHz datum 1.016 Kuhr et al. spectrum is unreliable below 160 MHz 0.906 Kuhr et al. spectrum may be high, misses 38 MHz datum 0.785 VLA flux density unestimated?; B-configuration data only 3C 452 142.5 142.6 ± 0.2 1.001 Note. — The uncertainties given for the VLA measurements are merely the formal standard deviations from fits to the images. Little significance should be attached to them. – 69 – Table 3. Zernike Polynomials Z l n (ρ, φ) n 1 2 3 0 l 1 2 3 · · · 2ρ2 − 1 · · · ρeiφ · · · (3ρ3 − 2ρ)eiφ · · · ρ2 e2iφ · · · · · · · · · ρ3 e3iφ Note. — The polynomials are expressed in terms of a radial distance from the phase cen- ter ρ and a position angle φ. The Z 0 0 polyno- mial is not used because it represents a total phase contribution (“piston”) to which the in- terferometer is insensitive. – 70 – Table 4. Evolving Techniques for Ionospheric Calibration Method Parameterization Imaging Capability FoV Baseline Reference Simple geometric shift None re-register position of known sources No restrictions Full field (λ/Dstation ) ≤5 km Erickson (1984) Classical φi (t) - one term Self Calibration per i station Bright, isolated ≤15′ sources (e.g. 3C ob jects) ≤400 km Kassim et al. (1993) Gizani et al. (2005) Lazio et al. (2006) Field-based calibration φ(t,α,δ) - single term for entire array Joint Multi-source Self-calibration φ(t,α,δ) - one term per i stations No restrictions (used for VLSS) Full field (λ/Dstation ) ≤12 km Cotton et al. (2004) No restrictions Full field (λ/Dstation ) ≤400 km TBD Table 5. 74 MHz VLA Classical Confusion Limits Configuration Resolution Confusion Level (′′ ) (mJy) A B C D 25 77 240 744 7 40 250 1500 – 71 – Table 6. 74 MHz VLA Polyhedral Imaging Configuration θfacet Nfacet (′ ) A B C D 13 22 39 69 720 250 80 25 Table 7. 74 MHz and 330 MHz Observations of Sources Source Alternative Integration Time 74 MHz 330 MHz Name A-conf. B-conf. C-conf. (min) (min) (min) 3C010 3C033 Perseus Cluster 3C084 3C098 3C129 3C144 3C218 3C219 3C274 3C327 3C353 3C390.3 3C392 3C405 3C445 3C452 3C461 Tycho SNR NGC1275 M1, Crab SNR Hydra A M87, Virgo A W44, SNR G34.7-0.4 Cyg A Cas A 71 41 51 51 44 13 92 16 82 71 52 51 72 51 9 31 76 7 80 62 73 73 53 93 14 16 82 71 51 51 81 51 9 42 82 9 35 29 111 111 36 87 19 18 58 49 21 27 27 27 10 · · · 33 7 beam (′′ ) 80 25 94 25 25 83×75 25 30 25 25 25 25 29×25 300 31×26 30 25 30 rms (Jy) 0.36 0.14 0.16 0.08 0.13 0.10 0.12 0.17 0.06 0.08 0.38 0.16 0.05 1.10 4.63 0.03 0.12 17.0 beam (′′ ) rms (mJy) 7.9×6.7 7.0 20.0 6.0 8.0 65.0 18.0 10.0 6.0 23.0 6.0 7.5 7.5 25.0×22 5.0 10.4×9.0 7.0 18.0 2.0 7.4 6.7 2.6 3.9 6.5 47.4 26.0 1.9 13.8 4.3 7.3 2.4 15.6 53.0 3.0 2.8 227.0 – 72 – Table 8. Peak Intensities and Flux Densities of 3C Sources at 74 MHz Name I (Jy beam−1) S (Jy) 3C 10 3C 33 3C 48 3C 84 3C 98 3C 123 3C 129 3C 144 3C 147 3C 161 3C 196 3C 218 3C 219 3C 273 3C 274 3C 286 3C 295 3C 298 3C 327 3C 353 13.0 55.3 66.9 82.2 35.7 387.8 27.8 305.6 57.2 85.6 133.1 270.2 38.9 142.2 567.2 27.6 111.6 95.3 52.8 152.2 3C 380 3C 390.3 3C 392 3C 405 3C 445 124.4 45.4 9.2 9308.3 9.2 252.1 ± 1.94 105.3 ± 0.41 67.6 ± 0.44 171.3 ± 0.52 98.0 ± 1.12 414.7 ± 1.31 70.4 ± 0.48 1811.3 ± 3.07 55.3 ± 0.82 87.5 ± 0.39 129.8 ± 0.95 644.2 ± 1.83 94.8 ± 0.50 140.6 ± 1.42 2084.6 ± 1.29 27.2 ± 0.76 107.9 ± 1.62 92.4 ± 1.22 118.4 ± 0.77 443.8 ± 0.76 143.7 ± 2.57 97.1 ± 0.42 715.7 ± 2.88 17205.0 ± 1.44 49.2 ± 0.32 – 73 – Table 8—Continued Name I (Jy beam−1 ) S (Jy) 3C 452 3C 461 3C 468.1 38.1 2362.2 42.0 142.6 ± 0.19 17693.9 ± 12.12 40.7 ± 0.66 This figure "f3.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f4.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f6a.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f6b.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f9.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f11.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f12.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f13.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f18.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20a.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20aa.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20ab.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20ac.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20ad.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20ae.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20af.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20ag.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20ah.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20ai.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20aj.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20b.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20c.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20d.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20e.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20f.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20g.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20h.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20i.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20j.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20k.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20l.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20m.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20n.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20o.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20p.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20q.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20r.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20s.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20t.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20u.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20v.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20w.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20x.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20y.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f20z.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1 This figure "f21.jpg" is available in "jpg"(cid:10) format from: http://arxiv.org/ps/0704.3088v1
astro-ph/0606417
1
0606
2006-06-16T21:17:22
Chandra Observations of Red Sloan Digital Sky Survey Quasars
[ "astro-ph" ]
We present short \chandra observations of twelve bright (i<18) z_em~1.5 quasars from the Sloan Digital Sky Survey chosen to have significantly redder optical colors than most quasars at the same redshift. Of the five quasars with optical properties most consistent with dust reddening at the quasar redshift, four show indirect evidence of moderate X-ray absorption (inferred N_H~10^22 cm^-2) with a dust-to-gas ratio <1% of the SMC value. The remaining seven objects show no evidence for even moderate intrinsic X-ray absorption. Thus, while optically red quasars are marginally more likely to show signatures of X-ray absorption than optically selected quasars with normal colors, dust-reddened type 1 AGN (as opposed to fully obscured type 2 AGN) are unlikely to contribute significantly to the remaining unresolved hard X-ray background. The red quasar population includes objects with intrinsically red continua as well as objects with dust-reddened continua. Improved sample selection is thus needed to increase our understanding of either subpopulation. To identify dust-reddened quasars likely to exhibit X-ray absorption, some measure of spectral curvature is preferable to simple cuts in observed or relative broad-band colors.
astro-ph
astro-ph
Draft version October 1, 2018 Preprint typeset using LATEX style emulateapj v. 11/12/01 6 0 0 2 n u J 6 1 1 v 7 1 4 6 0 6 0 / h p - o r t s a : v i X r a CHANDRA OBSERVATIONS OF RED SLOAN DIGITAL SKY SURVEY QUASARS Patrick B. Hall,1,2 S. C. Gallagher,3 Gordon T. Richards,1,4 D. M. Alexander,5 Scott F. Anderson,6 Franz Bauer,7 W. N. Brandt,8 Donald P. Schneider8 Draft version October 1, 2018 ABSTRACT We present short Chandra observations of twelve bright (i<18) zem∼1.5 quasars from the Sloan Digital Sky Survey chosen to have significantly redder optical colors than most quasars at the same redshift. Of the five quasars with optical properties most consistent with dust reddening at the quasar redshift, four show indirect evidence of moderate X-ray absorption (inferred NH ∼ 1022 cm−2) with a dust-to-gas ratio <1% of the SMC value. The remaining seven objects show no evidence for even moderate intrinsic X-ray absorption. Thus, while optically red quasars are marginally more likely to show signatures of X-ray absorption than optically selected quasars with normal colors, dust-reddened type 1 AGN (as opposed to fully obscured type 2 AGN) are unlikely to contribute significantly to the remaining unresolved hard X-ray background. The red quasar population includes objects with intrinsically red continua as well as objects with dust-reddened continua. Improved sample selection is thus needed to increase our understanding of either subpopulation. To identify dust-reddened quasars likely to exhibit X-ray absorption, some measure of spectral curvature is preferable to simple cuts in observed or relative broad-band colors. Subject headings: quasars: general -- X-rays: galaxies -- quasars: absorption lines 1. introduction It has long been suspected that optical surveys for ac- tive galactic nuclei are significantly incomplete due to orientation-dependent (and perhaps evolution-dependent) dust obscuration (Antonucci 1993). This suspicion has been confirmed by recent X-ray and infrared investigations that find higher densities of active galactic nuclei (AGN) than do optical surveys to comparable flux limits (e.g., Ueda et al. 2003; Bauer et al. 2004; Stern et al. 2005). The exact fraction of quasars missed by optical surveys is not yet clear, as the missing objects appear to be mostly lower-luminosity AGN (e.g., Ueda et al. 2003; Hao et al. 2005). Nonetheless, it is clear that a full understanding of quasars and their contribution to the X-ray background re- quires proper consideration of optically obscured quasars. Indeed, one of the most active topics in X-ray astronomy today is the origin of the hard X-ray background. With the sensitivity of Chandra and XMM-Newton, ∼90% of the X-ray background can be resolved into point sources at <6 keV, but only ∼60% at 6 -- 8 keV (e.g., Alexander et al. 2001; Worsley et al. 2005). The spectral shape of the > 10 keV X-ray background (Marshall et al. 1980) clearly indicates that a significant fraction of the very hard X-ray background comprises absorbed sources. The demograph- ics of this absorbed population are uncertain at present, and it is unclear how many are bona-fide type 2 AGN with fully obscured broad emission-line and optical continuum regions as opposed to X-ray-absorbed and optically dust- reddened but otherwise normal type 1 (broad line) AGN extincted beyond the flux limits of optical surveys. Characterizing this population requires not only statis- tical arguments demonstrating the existence of obscured quasars, but also focused investigations targeting the prop- erties of individual objects that exhibit evidence of obscu- ration. Some recent studies have claimed that quasars with red near-IR colors (J − K > 2) from the Two-Micron All-Sky Survey (2MASS) represent a population of pre- viously undiscovered quasars that could partially account for much of the "missing" hard X-ray background sources (Cutri et al. 2002; Wilkes et al. 2002, 2005). In this paper we take a complementary approach and investigate the na- ture of quasars that have relative UV/optical colors that are redder than average (at a given redshift). This selec- tion has the benefit of being less sensitive to color differ- ences due to emission lines moving in and out of the bands and of exhibiting larger differences due to extinction, since the effects of extinction are strongest at UV wavelengths. We use Chandra observations of these quasars to look for additional signs of extinction, such as a harder X-ray spec- trum and lower than expected X-ray flux. We present the construction of our sample in § 2, our X- ray observations and data analysis in § 3, the results of our observations in § 4, a discussion of our results in § 5, and our conclusions in § 6. We adopt H0 = 70 km s−1 Mpc−1, ΩΛ = 0.7, and Ωm = 0.3 (e.g., Spergel et al. 2006). 2. sample selection 2.1. Why Use Relative Optical Colors? 1 Princeton University Observatory, Princeton, NJ 08544. 2 Department of Physics & Astronomy, York University, 4700 Keele Street, Toronto, ON M3J 1P3, Canada. 3 Department of Physics & Astronomy, University of California, Los Angeles, 475 Portola Plaza, Mail Code 154705, Los Angeles, CA 90095-1547. 4 Department of Physics & Astronomy, Johns Hopkins University, 3400 N. Charles St., Baltimore, MD 21218. 5 Institute of Astronomy, Madingley Road, Cambridge CB3 0HA, UK. 6 Department of Astronomy, University of Washington, Box 351580, Seattle, WA 98195. 7 Department of Astronomy, Columbia University, 550 W. 120th Street, New York, NY 10027. 8 Department of Astronomy & Astrophysics, The Pennsylvania State University, University Park, PA 16802. 1 2 HALL ET AL. One of the primary problems with color-selecting a sam- ple of "red" quasars for further study is that quasar colors are a strong function of redshift. Large-area digital sur- veys such as the Sloan Digital Sky Survey (SDSS; York et al. 2000) are now enabling astronomers to character- ize the quasar color distribution as a function of redshift (Richards et al. 2001) and are, in fact, resolving their color distribution (Richards et al. 2003). While a fixed observed-frame color cut such as B − K > 5 does select quasars that are apparently red, as discussed by Richards et al. (2003) this approach fails to distinguish whether the quasar is intrinsically red (i.e., has a red [optically steep] power-law spectrum), has been reddened by dust extinction, or appears red because of strong emission or absorption in one of the bands or long-wavelength host galaxy contamination. For quasars of known redshift, it is possible to use SDSS colors alone to account for the red- shift dependence of quasar colors and, for sufficiently large reddening, to distinguish between a steeper power-law con- tinuum and the curvature of a dust-reddened continuum. Thus, to define a more redshift-independent red quasar sample, we take advantage of the well-defined structure in the color-redshift relation (see Fig. 1). For each of our quasars we can compute a relative color by subtracting the median color of quasars at that redshift (Richards et al. 2001). Relative colors are thus more nearly equally sensitive to reddened quasars at all redshifts and can be used to classify quasars as a function of color alone. For example, a quasar that has a relative u − g color of 0.2 (∆(u − g) = 0.2) is redder than the average quasar by 0.2 magnitudes in observed u − g regardless of its redshift. Such selection should yield a more homogeneous red AGN sample than a fixed observed-frame color cut such as J − K > 2 (Cutri et al. 2002). The relative color ∆(g − i) probes the shape of the UV/optical continuum due to the accretion disk at most redshifts studied by the SDSS, whereas a J −K selected sample measures the shape of the near-IR bump for z . 0.25 and the ratio of the near-IR to optical flux densities for 0.25 < z < 1. Further- more, although a J − K > 2 color cut does include dusty quasars, it also selects only slightly redder than average quasars with z < 0.5 (see Fig. 5 of Barkhouse & Hall 2001 and § 6.2 of Maddox & Hewett 2006). Near-IR selection of AGN is superior to optical selection when one is trying to avoid incompleteness caused by dust extincting objects below the survey flux limit. Yet, specif- ically because the UV/optical colors suffer more extinc- tion, UV/optical colors are better able to distinguish be- tween dust-reddened quasars and unreddened quasars with steeper than average continua. Our SDSS-based sample has one additional advantage in that the flux limit for the optically selected red quasars was determined in the i-band instead of the B-band, resulting in less of a bias against dust-reddened quasars (as compared to most other opti- cal surveys). Thus an investigation of red quasars using relative optical colors of SDSS quasars provides a happy medium for exploring the red quasar population: it has the benefits of relatively long wavelength selection, but covers wavelengths considerably affected by dust reddening. 2.2. Selection Details Our initial sample was selected from the 3814 Early Data Release (EDR) quasars (Schneider et al. 2002), giv- ing priority to bright objects (dereddened i < 18.0) with small Galactic NH. We considered only spatially unre- solved quasars with 0.6 ≤ zem ≤ 2.2 and ∆(g − i) > 0.2; see Figure 1. The color cut selects the reddest ∼11.5% of non-BAL, i < 18 quasars at each redshift, and the redshift selection avoids host galaxy contamination and reduces the effect of absorption from the Lyα forest. Avoiding resolved sources also reduces the chances that the red colors are due to host galaxy contamination. The sample was vetted to exclude quasars with broad absorption line (BAL) troughs (Reichard et al. 2003b) which are known to cause both optical reddening and X-ray absorption. (e.g., Gallagher et al. 2006). This was not possible for the eight zem < 1.6 quasars where C IV λ1549, the transition that most often exhibits a BAL trough, is not in the SDSS spectroscopic bandpass. Based on a raw high-ionization SDSS BAL frac- tion of 9.85% (Trump et al. 2006), we estimate there is a 38% chance that exactly one of these eight quasars is a high-ionization BAL quasar, a 15% chance that exactly two are, and a 3% chance that three or more are. We also excluded quasars detected in the 20 cm FIRST sur- vey (Becker, White, & Helfand 1995), since radio-powerful quasars have different X-ray properties (e.g., Shastri et al. 1993). Table 1 lists the 12 targets for which we obtained X-ray data. Figure 2 shows their SDSS spectra. An obvious question to ask is whether our selection criterion recovers quasars selected by J − K > 2 (and vice versa). In Figure 3 we show the relative g − i color versus the relative J − K color for SDSS-DR3 quasars also detected by 2MASS (Schneider et al. 2005). Ob- jects consistent with being dust reddened form a tail with red ∆(g − i) colors but undistinguished ∆(J − K) colors. In terms of observed J − K color we find that while our ∆(g − i) > 0.2 criterion does indeed recover some quasars with J − K > 2, only 1/3 of EDR quasars with J − K > 2 also have ∆(g−i) > 0.2. Since the tail of red ∆(g−i) colors comes predominantly from dust reddening (Hopkins et al. 2004), J − K > 2 selection must miss some dust-reddened quasars and include some that are not dust reddened. For example, none of our twelve targets have J − K > 2.9 We discuss J − K > 2 selection further in § 5.1. Note that our selection criteria were designed to select quasars that suffer from extinction (and reddening) by dust grains, not ones that simply have intrinsically red- der than average power-law continua. Unfortunately, the range of power-law slopes in quasars is such that it is not possible to do this with complete certainty. However, any unreddened but intrinsically red quasars that 'contami- nate' our sample are still of interest, since the relationship between the X-ray properties of quasars and their opti- cal/UV colors is not well characterized. 2.3. Damped Lyα Absorbers In our selection of targets, we did not reject objects with strong intervening absorption lines in their optical spectra. At first glance not rejecting such systems seems reasonable given the amount of reddening expected from a typical in- tervening absorption system. As shown by Richards et al. (2004), in a statistical sense (2003) and Hopkins et al. 9 For four of our targets without 2MASS J or K detections, we performed aperture photometry on calibrated 2MASS Atlas images. X-RAYING RED QUASARS 3 the reddening observed in SDSS quasars is dominated by SMC-like dust at the quasar redshifts. However, quasars with intervening absorption lines are three times as likely to have highly reddened spectra as quasars without them (York et al. 2006). And since SMC-like dust has a nearly featureless, nearly power-law exinction curve, it is not al- ways possible for a specific quasar to determine whether the dust is intrinsic to the quasar or along the line of sight. We can estimate the amount of optical reddening that is expected under the hypothesis that the galaxies causing absorption lines in some of our quasars are the source of reddening. Following Rao & Turnshek (2000, Fig. 25), if the rest-frame equivalent widths are EWMgII2796 > 0.5 A and EWFeII2600 > 0.5 A, there is a 50% chance the Mg II system comes from a DLA with NHI > 2 × 1020 cm−2, which gives E(B − V ) > 0.005 using the SMC extinction law. Quasars with such candidate DLAs are noted in Ta- ble 1. The two which also show Ca II absorption lines, J1310+0108 and J1323−0021, will be specially marked as they are the most likely to be affected by intervening ex- tinction (Wild & Hewett 2005). Note that for the maximum NHI found for any DLA along a quasar sightline to date, 5 × 1021 cm−2, the SMC extinction would be only E(B − V ) = 0.11. However, given recent work demonstrating that intervening ab- sorbers sometimes can redden quasars (Khare et al. 2005), it may be that these particular systems have LMC or MW extinction laws which would require a multiplication of E(B − V ) by factors of 2 or 9, respectively. If that is the case, then the redness of such objects may not be intrin- sic and they cannot be used to address the issue of the X-ray properties of intrinsically red or reddened quasars. We consider this issue further in § 4.2. 3. x-ray observations and data analysis Chandra observed our 12 red quasar targets between 2003 November 09 and 2004 September 29 (see Table 2). Each target was observed at the aimpoint of the back- illuminated S3 CCD of the Advanced CCD Imaging Spec- trometer (ACIS; Garmire et al. 2003) in faint mode. In general, data analysis followed the procedure de- tailed in Gallagher et al. (2005), which we outline briefly. Both aperture photometry and the CIAO version 3.2 (see http://cxc.harvard.edu/ciao) wavelet detection tool wavdetect (Freeman et al. 2002) were used in the soft (0.5 -- 2 keV), hard (2 -- 8 keV), and full (0.5 -- 8 keV) bands to determine the measured counts for a point source in each band. Except for J1133+0058, the quasars were each significantly detected with full-band counts ranging from 7 to 237; the difference in counts between wavdetect and aperture photometry was always ≤ 3 counts. The back- ground was in all cases negligible (<1 count in the source region). At the median flux of our sample the source den- sity is 8×10−7 arcsec−2 (e.g., Bauer et al. 2004), so the chance of a misidentification is extremely small. For each target, to provide a coarse quantitative mea- sure of the spectral shape we calculated the hardness ra- tio HR=(h−s)/(h+s), where h and s refer to the hard- and soft-band counts, respectively. A typical, radio- quiet quasar has a power-law continuum in the 0.5 -- 10.0 keV band characterized by the photon index, Γ, and the 1 keV normalization, N1keV: fE = N1keVE−Γ photon cm−2 s−1 keV−1. From X-ray spectroscopic stud- ies, Γ is found to average 2.0 ± 0.25 for radio-quiet quasars (e.g., George et al. 2000; Reeves & Turner 2000); such an average quasar would be observed to have HR ≈ −0.63. To transform the observed HR into Γ, the X-ray spectral modeling tool XSPEC (Arnaud 1996) was used. The de- tector response to incident power-law spectra with varying Γ was simulated. The hardness ratio and errors were com- pared to the modeled HR to determine the ΓHR that would generate the observed HR. The uncertainties on ΓHR re- flect the statistical uncertainties in the HR. The modeled full-band count rate was normalized to the observed full- band count rate to obtain the power-law normalization, N1keV. Using N1keV and ΓHR, the 0.5 -- 8.0 keV flux, FX, and the flux density at rest-frame 2 keV, f2keV, were cal- culated. The errors quoted for these two values are the Poisson errors (Gehrels 1986) from the full-band counts. Lastly, we calculated the optical/UV to X-ray in- dex αox=log(f2keV/f2500)/ log(ν2keV/ν2500) and its uncer- tainty. The quantity f2500 is the average flux density within the rest-frame range 2500 ± 25 A in the SDSS spec- trum, scaled to account for flux outside the spectroscopic fiber aperture. We took the scale factor to be the flux ratio between the object's photometric PSF magnitude and its fiber magnitude as synthesized from the spectrum, with both magnitudes measured in the filter into which rest- frame 2500 A is redshifted. As detailed in §2.2 of Gallagher et al. 2005, the quoted errors in αox incorporate both the Poisson uncertainty in f2keV and the estimated effects of UV variability between the SDSS and Chandra observa- tions (Ivezi´c et al. 2004), which are on average 500 days apart in the quasar rest frame. We also calculated ∆αox= αox−αox(L2500); a quasar's ∆αox is its αox relative to the expected αox for a quasar of its luminosity (Strateva et al. 2005, Eq. 6; see also Steffen et al. 2006). These quantities are presented in Table 3. 3.1. X-ray Spectral Fitting For those six quasars with > 50 full-band counts, we also performed X-ray spectral fitting. The spectra were ex- tracted using 2.′′5-radius source cells. Our spectral model was a simple absorbed power-law continuum with both Galactic and intrinsic absorption by neutral gas with solar abundances at the quasar redshift; the photon statistics do not warrant a more complex model. In this low-count regime, the spectra are not binned and the fit is performed with XSPEC by minimizing the C-statistic (Cash 1979). The best-fitting Γ and intrinsic NH for the six quasars with > 50 counts are listed in Table 4. Since these quasars have the highest X-ray fluxes in our sample, they are by defini- tion unlikely to be absorbed in the X-ray; indeed, for only one of them is there a significant detection of absorption. This object, J1323−0021, exhibits the most metal-rich in- tervening absorption system currently known (§4.2.7). 4. results 4.1. Correlations Between X-ray and Optical Properties We begin our X-ray analysis by searching for evidence of X-ray absorption that may occur in conjunction with optical dust extinction and reddening. The signatures of absorption in snapshot X-ray observations are flatter than average X-ray photon indices (i.e., Γ . 1.5), since 4 HALL ET AL. absorption preferentially removes softer X-rays, and un- usual values of the optical to X-ray flux ratio, αox, or the luminosity-corrected optical to X-ray flux ratio, ∆αox. We first examine the relationship between ΓHR and ∆(g − i), which is shown in Figure 4. Richards et al. (2003) showed that most quasars have ∆(g − i) between −0.2 and 0.2 and that there exists a tail of redder quasars consistent with being dust reddened. The typical range of Γ is found to be 2.0 ± 0.25 for radio-quiet quasars (e.g., George et al. 2000; Reeves & Turner 2000). Dust causes quasars to appear redder in the optical, while gas absorbs soft X-ray photons. For any observed dust reddening, the accompanying X-ray absorption will be greater (and thus the ΓHR lower) if the dust-to-gas ratio is smaller. Quasars in the lower right hand corner of panel (a) in Figure 4 are potentially obscured quasars. In panel (b) of Figure 4 we show ΓHR vs. the ultravi- olet spectral slope, αUV (Fν ∝ ναUV ). αUV measures the quasar's underlying continuum slope at fixed rest-frame wavelengths that are relatively devoid of emission line flux and is thus a more precise measurement of the continuum color than ∆(g − i). Our new data appear to reaffirm the correlation between ΓHR and αUV found by Gallagher et al. (2005). Combining both datasets, the chance of no correlation is only 0.03% using Kendall's τ test. Even considering only quasars with ∆αox ≥ −0.1, the chance of no correlation is still only 0.95%. In the plot we show the BCES(YX) (Akritas & Bershady 1996) linear fit to the full combined datasets, ΓHR = (0.54±0.15)αUV +(2.07±0.13). More work, however, is needed to disentangle a correlation between ΓHR and intrinsic UV spectral slope from a corre- lation induced by the effects of dusty gas on both parame- ters. That is, an intrinsically red quasar might have an in- trinsically hard X-ray spectrum, whereas a dust-reddened quasar might have an intrinsically soft X-ray spectrum that appears hard due to X-ray absorption. We next search for for any trends between color and ∆αox as shown in Figure 5. In panel (a) the lines show the expected relationship between ∆αox and ∆(g − i) for neutral gas with three different SMC-like dust-to-gas ra- tios. For an SMC dust-to-gas ratio, even minimal X-ray absorption can be accompanied by heavy extinction of op- tical light by dust, and the ratio of emergent X-ray to optical fluxes (in other words, ∆αox) increases. But as the dust-to-gas ratio decreases, X-ray wavelengths become in- creasingly affected relative to the optical. The open points are roughly consistent with a typical SMC reddening as shown by the solid line. The filled points, however, may require a different explanation. Both panels of Figure 5 show that there is intrinsic scatter in ∆αox which is inde- pendent of a quasar's optical/UV color, but that quasars with relatively red colors may be more likely to be X- ray faint than are quasars with relatively blue colors.10 When αUV is used as a measure of optical color to split the full plotted sample in half at αUV=−0.75, an invariant Kolmogorov-Smirnov (Kuiper) test shows that the ∆αox distributions for the two halves are different at the 89.2% significance level. However, when the sample is split in half at ∆(g − i)=0.26, there is only a 55.8% chance the distributions are different. To test whether the extrema in these figures are consis- tent with absorption in the X-ray, in Figure 6 we show ΓHR vs. ∆αox overlaid on curves that demonstrate the effects of X-ray absorption on a quasar of intrinsic ΓHR=2 and ∆αox=0. The heavy curves are for gas with various dust- to-gas ratios at a fixed z = 1.5, while the dotted curves are for completely dust-free neutral gas at the redshifts shown. Dust-free gas absorbs soft X-rays, but does not affect the optical spectrum until the absorption becomes Compton- thick. Thus, the effect of such gas in Figure 6 is to decrease both ∆αox and ΓHR. The three points closest to the lower left-hand corner appear consistent with the absorption hy- pothesis and inconsistent with being due to intrinsic scat- ter in ΓHR and ∆αox. Furthermore, the distribution of points in the diagram is not consistent with random scat- ter around ΓHR=2 and ∆αox=0, but is consistent with quasars with intrinsic ΓHR=2 and ∆αox=0 being affected by absorption of varying dust-to-gas ratios, as indicated by the various tracks, HR=2 and ∆αox=0. However, with the typical measurement accuracy and probable intrinsic scat- ter of ΓHR and ∆αox, this Figure by itself can only be used to unambiguously identify moderately reddened quasars if they have low dust-to-gas ratios. Moderate reddening with a high dust-to-gas ratio can produce a marked effect on a quasar's optical spectrum without substantially affecting its X-ray properties; a quasar's optical/UV spectrum is an invaluable source of information on the reddening of such quasars. (See § 4.2 and § 5 for specific examples.) 4.2. Notes on Individual Quasars Before we can explore the X-ray properties of dust- reddened quasars as a population, we must examine the diagnostic information in Figures 2, 4, and 5 to determine which objects in our sample are most likely to be reddened by dust versus intrinsically red. Both of these groups could suffer from X-ray absorption, which does not always occur in conjunction with dust extinction (e.g. Brandt, Laor, & Wills 2000). The shape of the optical/UV continuum, the relationship between the relative colors, and the X- ray properties affect this determination. We now examine these parameters for each of our objects. 4.2.1. J0002+0049; zem = 1.355, ∆(g − i) = 0.396, αUV = −0.97, ΓHR = 2.13, ∆αox = 0.11 The optical spectrum shows moderate reddening, for which the intervening absorption system is likely too weak to be responsible. This quasar has the softest X-ray spec- trum in the sample and ∆αox is positive, thus there is no sign of X-ray absorption; the results from spectral- fitting are consistent with this conclusion. These prop- erties might lead us to believe that this quasar is intrinsi- cally red. However, there is evidence for curvature in the optical spectral energy distribution (SED) from the SDSS photometry. Specifically ∆(u − r) is larger than ∆(g − i), which is larger than ∆(r − z), as expected for a dust- reddened object. (For an object with a power-law contin- uum and unexceptional emission-line properties ∆(u − r), ∆(g − i), and ∆(r − z) would be equal within the uncer- tainties.) Thus we conclude that the object is reddened by SMC dust with a normal dust-to-gas ratio; this leads to noticable curvature of the optical spectrum but negligibly effects the X-ray spectrum. 10 The converse, that relatively X-ray faint quasars are more likely to be abnormally red, is also probably true. X-RAYING RED QUASARS 5 4.2.2. J1133+0058; zem = 1.937, ∆(g − i) = 0.382, αUV = −1.15, ΓHR unknown, ∆αox =< −0.56 The optical spectrum is consistent with mild dust red- dening rather than a flatter than typical power-law slope. Dust reddening is also indicated by the fact that ∆(u − r) > ∆(g − i) > ∆(r − z). There is a strong intervening absorption system (a probable DLA) at z = 1.786. This quasar was undetected in our Chandra observation and must thus be very heavily absorbed in the X-ray (equiv- alent NH & 7 × 1022 cm−2 at z ≃ 1.9). A DLA could provide such a column if it was dominated by H2 (Schaye 2001) but nonetheless had a low dust-to-gas ratio, which is rather contrived, or if had a metallicity ∼10 times solar, which is rather unlikely. We conclude that this quasar ex- periences some intrinsic X-ray absorption and some from the intervening system, and similarly for its dust redden- ing. (There is moderately strong narrow C IV absorption 7000 km s−1 from the quasar redshift, which could be a tracer of high-ionization gas intrinsic to the quasar.) We discuss this issue further in §4.2.6 and §4.2.7. In any case, the combination of moderate dust reddening and strong X-ray absorption means that the overall dust-to-gas ratio along the line of sight to this quasar is rather small. 4.2.3. J1226−0011; zem = 1.176, ∆(g − i) = 0.261, αUV = −0.72, ΓHR = 2.01, ∆αox = 0.01 The optical spectrum is only mildly red, but dust extinc- tion might be present since ∆(u−r) > ∆(g −i) > ∆(r−z). There is no indication of X-ray absorption. This quasar could be either intrinsically red or mildly dust reddened. 4.2.4. J1251+0002; zem = 0.878, ∆(g − i) = 0.319, αUV = −1.01, ΓHR = 1.24, ∆αox = −0.25 The spectrum of this quasar has little rest-frame UV coverage, but the evidence of moderate curvature in the spectrum and photometry [∆(u−r) > ∆(g −i) > ∆(r −z)] are consistent with dust reddening. The values of ∆αox and ΓHR imply significant X-ray absorption. We conclude that this quasar is dust-reddened with X-ray absorption. 4.2.5. J1302+0000; zem = 1.797, ∆(g − i) = 0.363, αUV = −0.86, ΓHR = 1.52, ∆αox = −0.16 The optical spectrum is normal. While there is interven- ing absorption, there is no indication of spectral curvature from reddening. The values of ∆αox and ΓHR are consis- tent with moderate X-ray absorption, but are only ∼ 2σ from the typical values seen in quasars without X-ray ab- sorption. We argue that the redness of this object is most likely intrinsic and not due to dust. 4.2.6. J1310+0108; zem = 1.392, ∆(g − i) = 0.626, αUV = −1.61, ΓHR = 0.84, ∆αox = −0.42 The optical spectrum of this quasar is heavily red- dened. There is strong intervening absorption (a probable DLA) at z = 0.8621 and Ca II absorption, and accord- ing to (Wild & Hewett (2005) the quasar is reddened by E(B − V ) = 0.209 assuming an LMC extinction curve. However, this quasar is significantly absorbed in the X- ray, consistent with an equivalent solar-metallicity absorb- ing column of NH ≃ (5 ± 2) × 1022 cm−2. This is the same situation as for J1133+0058, and we reach the same conclusion: this quasar has both intrinsic and interven- ing X-ray absorption, and similarly for its dust reddening. Unfortunately, the quasar is at too low a redshift to look for intrinsic C IV absorption as a possible tracer of high- ionization gas. We discuss this issue further in §4.2.7. 4.2.7. J1323−0021; zem = 1.388, ∆(g − i) = 0.463, αUV = −1.12, ΓHR = 1.63, ∆αox = 0.05 The optical spectrum of this quasar is heavily reddened, has strong intervening absorption estimated to be of DLA strength (see §2.3), and has Ca II absorption that could be accompanied by optical reddening. (This intervening sys- tem is at too low a redshift to be in the (Wild & Hewett (2005) sample.) This quasar has unremarkable ΓHR and ∆αox values, but X-ray spectral-fitting yields Γ = 2.1±0.4 and a significant detection of absorption. If the X-ray ab- sorption is set to the DLA redshift of z = 0.716, then NH = 5.7+3.7 −3.1 × 1021 cm−2 (68% confidence for 2 parame- ters of interest); an intrinsic absorber requires more than double that column density (see Table 4). Our absorption detection is consistent with the re- (2006) that this is the most sult of Peroux et al. metal-rich intervening quasar absorption system known, with [Zn/H]=+0.61±0.20 (metallicity 4.1+2.4 −1.5 times solar). However, even after adjusting our best-fit solar-metallicity X-ray column downward by a factor of four, the resulting NH measured in the X-ray is still a factor of 9+8 −6 larger than this system's neutral NHI= 1.62+1.01 −0.55×1020 cm−2 measured in the UV (Khare et al. 2004). We conclude that an intervening absorption system causes the optical reddening toward this quasar and some of the X-ray absorption toward it, but not all: we infer that there is also high-ionization, X-ray-absorbing gas in- trinsic to this quasar. In fact, the spectrum of this quasar shows two strong, narrow C IV absorption systems very near the quasar redshift (Khare et al. 2004), at least one of which is accompanied by strong, narrow N V absorption in an HST spectrum (GO project 9382; PI: S. Rao). It is somewhat suspicious to postulate that there are three quasars in our sample which have both dust redden- ing and strong intervening absorption systems but which also require intrinsic X-ray absorption at the quasar red- shift (this quasar, J1133+0058 and J1310+0108), even if there is possible evidence for intrinsic high-ionization gas in the first two cases. Nonetheless, this conclusion appears more likely than the alternative that the intervening ab- sorbers are as strong as any ever seen along quasar sight- lines and have metallicities ∼10 times solar. This is espe- cially true for J1323−0021, where we know the absorber's metallicity and NHI and there is still a factor of 10 discrep- ancy with the X-ray column. Moreover, given that dusty intrinsic and intervening absorption systems do exist, both types of systems will be overrepresented in our red quasar sample, and rare systems where both effects are at work will be even more overrepresented. In any case, our con- clusion can be tested with better X-ray data to constrain the X-ray absorbing columns (and to ensure that the mea- sured X-ray columns are not systematically high) and with high-resolution UV spectroscopy to constrain the NHI and metallicities of the three intervening absorption systems. 6 HALL ET AL. 4.2.8. J1708+6154; zem = 1.415, ∆(g − i) = 0.259, αUV = −1.13, ΓHR = 1.58, ∆αox = 0.13 This quasar's optical spectrum is moderately red. It has ∆(u − r) ≃ ∆(g − i) > ∆(r − z). The X-ray properties are within the normal range in ∆αox and ΓHR. All the above evidence is consistent with this quasar being either intrinsically red or only mildly dust-reddened. 4.2.9. J1714+6119; zem = 1.847, ∆(g − i) = 0.426, αUV = −0.97, ΓHR = 1.32, ∆αox = −0.10 In this quasar, the optical spectrum is mildly red. The X-ray spectrum is fairly hard, but ∆αox within the normal range. Given that both the C IV and He II emission lines in this quasar are strong and narrow, and that it has red colors in ∆(r − z) and ∆(g − i) but a blue color in ∆(u − r) (due to strong Lyα emission?), this object would appear to be intrinsically red rather than dust-reddened. (Richards et al. (2003) presented evidence that quasars with high equivalent width emission lines tend to be redder.) 4.2.10. J1715+6323; zem = 2.182, ∆(g − i) = 0.352, αUV = −1.05, ΓHR = 1.75, ∆αox = 0.02 The optical spectrum shows mild reddening. The object has red ∆(r − z) and ∆(g − i) but blue ∆(u − r). There is no strong evidence for X-ray absorption from spectral fitting. This object is most likely to be intrinsically red. 4.2.11. J1735+5355; zem = 0.956, ∆(g − i) = 0.235, αUV = −0.25, ΓHR = 1.73, ∆αox = 0.07 In this quasar, which has the bluest ∆(g −i) value in our sample, the optical spectrum is normal or somewhat blue (note the quasar's much bluer position in Figure 4b rela- tive to the rest of our sample). The Fe II emission lines are very weak; this might be affecting the ∆(g − i) color. This quasar would appear to be neither dust reddened or in- trinsically red, but it does have an abnormal relative color which has caused it to 'contaminate' our sample. 4.2.12. J1738+5837; zem = 1.279, ∆(g − i) = 0.320, αUV = −0.70, ΓHR > 0.48, ∆αox = −0.57 Our final quasar has an optical spectrum that is mildly reddened. It also has ∆(u − r) > ∆(g − i) > ∆(r − z), consistent with dust reddening. The ∆αox and ΓHR values indicate significantly X-ray absorption. Thus we conclude that this object is dust-reddened as well as X-ray absorbed. 4.2.13. Summary of Notes on Individual Quasars The quasars in our relatively red quasar sample appear to have a variety of dominant causes for their red colors: Intrinsically red: 3 (§4.2.5, §4.2.9, §4.2.10) Dust-reddened: 3-5 (§4.2.1, §4.2.2?, §4.2.4, §4.2.6?, §4.2.12) Either: 2 (§4.2.3, §4.2.8) Reddened by intervening absorber: 1-3 (§4.2.2?, §4.2.6?, §4.2.7) Misclassified as red: 1 (§4.2.11) 5. discussion necessarily X-ray absorbed. However, the greater range of optical/UV to X-ray flux ratios and X-ray hardness ratios in quasars that appear optically red shows that they have a (marginally significant) tendency to be X-ray absorbed more often than optically blue quasars. Unfortunately, it does not appear straightforward to accurately predict which red quasars will be absorbed in the X-ray using the optical/UV colors alone. In our small sample of twelve relatively red quasars, five show some evidence for intrinsic X-ray absorption based on the combination of both more negative ∆αox values and lower ΓHR values (when we can constrain ΓHR) than nor- mal quasars. It is plausible that at least two effects are at work: 1) Intrinsically red quasars have intrinsically harder X-ray spectra but normal ∆αox values (Gallagher et al. 2005); 2) Dust-reddened quasars have X-ray spectra hard- ened (though not necessarily significantly) by absorption in gas accompanying the observed dust, and ∆αox values decreased by X-ray absorption (for low dust-to-gas ratios). Interpreting the distributions of ∆αox and ΓHR for rel- atively red quasars can be further complicated by correla- tions of X-ray and dust reddening properties of quasars with other quasar properties. Gallagher et al. (2005) find that quasars with large C IV blueshifts have X-ray absorption without significant dust reddening, since such quasars have relatively blue optical/UV spectra on average (Richards et al. 2002). Reichard et al. (2003a) noted that the heavily X-ray-absorbed broad absorption line quasars are dust-reddened but intrinsically blue. Thus, even with exploratory X-ray data it can be diffi- cult to securely distinguish intrinsically red quasars from quasars reddened by SMC-like dust. To demonstrate this, in Figure 7 we replot ΓHR vs. ∆αox. We have added an outline of the region in which z ∼ 2 broad absorption line (BAL) quasars are found (Gallagher et al. 2006). We have also added points for quasars with ΓHR and ∆αox readily available in the literature; namely, the unusual, optically very blue and luminous, yet X-ray faint quasar PHL 1811 (Leighly, Halpern, & Jenkins 2004; Choi, Leighly, & Mat- sumoto 2005) and the putative soft X-ray weak AGN of Risaliti et al. (2003) objects with the most negative ∆αox also have the lowest ΓHR (the quasars in question are HS 0848+1119 and HS 0854+0915, discounting the BAL quasar HS 1415+2701 which has only an upper limit on ∆αox). Yet even though both fall on the neutral absorption tracks, the lack of spec- tral curvature in their spectra points to their being intrin- sically red rather than dust reddened (although there is a ≃10% chance that HS 0854+0915 is an unrecognized high-ionization BAL quasar). (2003).11 The two Risaliti et al. Similarly, our best estimate is that the two uppermost filled squares in Figure 6 are quasars with intrinsically red SEDs, but these two objects also lie on the tracks of neu- tral absorption. Thus, falling on the neutral absorption track in this diagram is no guarantee that neutral absorp- tion is causing the X-ray faintness and spectral flatness. X-ray spectral fitting may help resolve such a degeneracy, but only if the spectrum has sufficient counts to be fit. The data presented here show that quasars selected to have the reddest optical colors at a given redshift are not 11 As pointed out by Brandt, Schneider, & Vignali (2004) and Strateva et al. (2005), the bulk of the Risaliti et al. (2003) objects are not actually X-ray weak if the luminosity dependence of αox is taken into account. Possibly the best candidate for an intrinsically red quasar in our Chandra sample is J0156+0053 from Gal- X-RAYING RED QUASARS 7 lagher et al. (2005). This object has ΓHR≃1, ∆αox≃0, and an optical/UV spectrum with a red continuum and strong emission lines. However, that classification could not be made without reference to the SDSS spectrum be- cause in all our plots this object appears consistent with dust reddening (specifically, E(B − V ) ≃ 0.08 with a dust- to-gas ratio 10% of the SMC ratio). Finally, note that the three leftmost quasars in Figure 6 have properties marginally inconsistent with X-ray absorp- tion from gas with a dust-to-gas ratio even just 10% as large as observed in the SMC, as might be expected from gas in an intervening galaxy. Instead, they are most consis- tent with absorption from large columns of partially ion- ized gas with dust-to-gas ratios <1% of the SMC value, consistent with the expected characteristics of gas in the immediate quasar environment. These X-ray-faint objects do overlap part of the ΓHR-∆αox diagram inhabited by z ∼ 2 BAL quasars (Gallagher et al. 2006), but lie much closer to the neutral absorption tracks for their redshifts than do most BAL quasars. These X-ray-faint objects are unlikely to be BAL quasars but could nonetheless have either ionized intrinsic absorption, or neutral intrinsic ab- sorption with 90 -- 95% partial covering. 5.1. Comparison with J − K > 2 AGN Though both samples are referred to as red, our sample and the J − K > 2 AGN studied in the X-ray by Wilkes et al. (2005) are substantially different. The primary dif- ferences are that the J − K > 2 AGN are less luminous and at lower redshift. Both differences make it more diffi- cult in J − K > 2 samples to isolate the nuclear emission in the optical/near-IR (from the host galaxy's flux) and in the X-ray (from AGN- or starburst-related soft excesses and a larger fraction of hard-band flux from reflection). For example, substantial reddening of J − K > 2 AGN has been claimed based on the low I-band luminosities of many J −K > 2 AGN in HST imaging (Marble et al. 2003) and low R-band luminosities from CFHT imaging (Hutch- ings et al. 2003). An equally plausible interpretation is that at least some J − K > 2 objects are low-luminosity AGN in luminous host galaxies with substantial near-IR emission from hot nuclear dust, which can produce quite red J − K colors if present in sufficient quantities (Scoville et al. 2000; Rodr´ıguez-Ardila & Mazzalay 2006). If correct, this interpretation would also help explain the X-ray faintness of the four J − K > 2 AGN in Wilkes et al. (2005) which are not type 2 AGN: they would in fact have normal ∆αox. (Note that only one of the four has a value of Γ(2-10 keV) harder than the range spanned by our quasar sample, but that they lack ∆αox values because their in- trinsic nuclear L2500 are not known.) To understand the intrinsic nuclear luminosities of J − K > 2 AGN, high- resolution near-IR imaging is needed to directly measure the relative AGN and host galaxy contributions to the in- tegrated J and K band magnitudes. While quasar samples selected to have J − K > 2 do in- clude red quasars,12 the lack of a relation between ∆(g −i) This work was enabled by Chandra X-ray Center grant and ∆(J − K) in Figure 3 suggests that J − K > 2 se- G04 -- 5116X. PBH acknowledges support from NSERC, lection is not the optimal way to select the most complete SCG from NASA through the Spitzer Fellowship Program and least contaminated samples of dust-reddened AGN. under award 1256317, DMA from the Royal Society, WNB Of course, our relatively red quasar sample will miss the 12 For example, a high polarization subsample chosen for spectropolarimetry has revealed some reddened broad-line AGN (Smith et al. 2003). most heavily reddened quasars and has its share of con- taminants as well. Therefore, better samples are needed to understand the dust reddened quasar population and to separate it from the intrinsically red quasar popula- tion. The most promising additional criterion to use in se- lecting dust-reddened quasars is spectral curvature, which arises from the extinction properties of small dust grains. Spectral curvature should be measurable from broadband photometry (e.g., §4.2.1), especially if it extends from the rest-frame far-UV (e.g., GALEX) through the near-IR, as well as from well-calibrated, wide-λ-coverage spectroscopy. 6. conclusions We have used short Chandra observations to demon- strate that optically red quasars are marginally more likely to exhibit evidence for X-ray absorption than optically blue quasars, but that not all optically red quasars ex- hibit X-ray absorption (§ 4 and § 5). Four of the five quasars in our sample with optical spectra most consis- tent with dust reddening at the quasar redshift (§ 4.2.13) show signatures in ∆αox and ΓHR of X-ray absorption from neutral or partially ionized gas with column den- sities NH & 2 × 1022 cm−2 and dust-to-gas ratios <1% of the SMC value (§ 4.1 and § 5). However, half of our sample shows no evidence for such absorption. We conclude that dust-reddened type 1 AGN (as opposed to fully obscured type 2 AGN) are unlikely to contribute significantly to the remaining unresolved hard X-ray background, at least for the levels of extinction probed by our sample. Plausible origins have been identified (see the Appendix) for most of the intrinsic dispersion exhibited by quasars in αox (the correlation of αox with luminosity, plus flux vari- ability in both bands) and possibly in ΓHR (the possible correlation of ΓHR with αUV, plus X-ray spectral variabil- ity). However, the existence of rare outliers like PHL 1811 underscores our ignorance of how those parameters inter- act to produce the observed αox and ΓHR values. Progress in understanding the X-ray properties of the type 1 radio-quiet quasar population will require even more careful work to create samples that are as homo- geneous as possible (§ 5.1). Such samples must not only be thoroughly vetted to remove quasars exhibiting strong intervening absorption and BAL troughs, but should also span the observed ranges of luminosity, C IV blueshift, αUV , optical/UV spectral curvature, estimated MBH , and possibly emission-line EW, since all those properties may be interrelated with a quasar's X-ray properties. While such studies can make use of serendipitous archival obser- vations of random quasars (e.g., Gallagher et al. 2005; Strateva et al. 2005; Risaliti & Elvis 2005), the surface density of relatively bright quasars on the sky is low and the the fraction of the sky covered by hard X-ray obser- vations is small. Serendipitous observations do not yield samples that cleanly span the full range of quasar proper- ties, and so there is a continuing need for targeted X-ray observations of carefully selected quasar samples. 8 HALL ET AL. from NASA LTSA grant NAG5 -- 13035, and DPS from NSF grant AST03 -- 07582. Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Founda- tion, the U.S. Department of Energy, the National Aero- nautics and Space Administration, the Japanese Monbuk- agakusho, the Max Planck Society, and the Higher Edu- cation Funding Council for England. The SDSS Web Site is http://www.sdss.org/. REFERENCES Akritas, M. G. & Bershady, M. A. 1996, ApJ, 470, 706 Alexander, D., Brandt, W., Hornschemeier, A., et al. 2001, AJ, 122, 2156 Antonucci, R. 1993, ARA&A, 31, 473 Arnaud, K. A. 1996, in ADASS V, ed. G. Jacoby & J. Barnes, 17 Barkhouse, W. A. & Hall, P. B. 2001, AJ, 121, 2843 Bauer, F. E., Alexander, D. M., Brandt, W. N., et al. 2004, AJ, 128, 2048 Bauer, F. E., Alexander, D. M., N., B. W., et al. 2004, AJ, 128, 2048 Becker, R. H., White, R. L., & Helfand, D. J. 1995, ApJ, 450, 559 Brandt, W., Schneider, D., & Vignali, C. 2004, in AGN Physics with the Sloan Digital Sky Survey, ed. G. T. Richards & P. B. Hall, 303 Brandt, W. N., Laor, A., & Wills, B. J. 2000, ApJ, 528, 637 Cash, W. 1979, ApJ, 228, 939 Choi, J., Leighly, K. M., & Matsumoto, C. 2005, American Astronomical Society Meeting Abstracts, 206 Cutri, R. M., Nelson, B. O., Francis, P. J., et al. 2002, in AGN Surveys, ed. R. F. Green, E. Y. Khachikian, & D. B. Sanders, 127 Dickey, J. M. & Lockman, F. J. 1990, ARA&A, 28, 215 Freeman, P. E., Kashyap, V., Rosner, R., et al. 2002, ApJS, 138, 185 Gallagher, S., Richards, G., Hall, P., et al. 2005, AJ, 129, 567 Gallagher, S. C., Brandt, W. N., Chartas, G., et al. 2006, ApJ, in press (astro-ph/0404487) Garmire, G. P., Bautz, M. W., Ford, P. G., et al. 2003, in X-Ray and Gamma-Ray Telescopes and Instruments for Astronomy, Proc. SPIE 4851, ed. J. E. Truemper & H. D. Tananbaum, 28 Gehrels, N. 1986, ApJ, 303, 336 George, I. M., Turner, T. J., Yaqoob, T., et al. 2000, ApJ, 531, 52 Hao, L., Strauss, M. A., Fan, X., et al. 2005, AJ, 129, 1795 Hopkins, P. F., Strauss, M. A., Hall, P. B., et al. 2004, AJ, 128, 1112 Hutchings, J. B., Maddox, N., Cutri, R. M., et al. 2003, AJ, 126, 63 Ivezi´c, Z., Lupton, R. H., Juric, M., et al. 2004, in The Interplay Among Black Holes, Stars and ISM in Galactic Nuclei, ed. T. Storchi Bergmann, L. C. Ho, & H. R. Schmitt, 525 Khare, P., Kulkarni, V., Lauroesch, J. T., et al. 2004, ApJ, 616, 86 Khare, P., York, D. G., vanden Berk, D., et al. 2005, in IAU Colloq. 199: Probing Galaxies through Quasar Absorption Lines, ed. P. Williams, C.-G. Shu, & B. Menard, 427 -- 429 Kraft, R. P., Burrows, D. N., & Nousek, J. A. 1991, ApJ, 374, 344 Laor, A., Fiore, F., Elvis, M., et al. 1994, ApJ, 435, 611 Leighly, K., Halpern, J., & Jenkins, E. 2004, in AGN Physics with the Sloan Digital Sky Survey, ed. G. T. Richards & P. B. Hall, 277 Lyons, L. 1991, Data Analysis for Physical Science Students (Cambridge: Cambridge University Press) Maddox, N. & Hewett, P. C. 2006, MNRAS, 176 Manners, J., Almaini, O., & Lawrence, A. 2002, MNRAS, 330, 390 Marble, A., Hines, D., Schmidt, G., et al. 2003, ApJ, 590, 707 Markowitz, A., Edelson, R., & Vaughan, S. 2003, ApJ, 598, 935 Marshall, F. E., Boldt, E. A., Holt, S. S., et al. 1980, ApJ, 235, 4 Mateos, S., Barcons, X., Carrera, F. J., et al. 2005, A&A, 444, 79 Peroux, C., Kulkarni, V., Meiring, J., et al. 2006, A&A Rao, S. M. & Turnshek, D. A. 2000, ApJS, 130, 1 Reeves, J. N. & Turner, M. J. L. 2000, MNRAS, 316, 234 Reichard, T., Richards, G., Hall, P., et al. 2003a, AJ, 126, 2594 Reichard, T., Richards, G., Schneider, D., et al. 2003b, AJ, 125, 1711 Richards, G., Hall, P., Vanden Berk, D., et al. 2003, AJ, 126, 1131 Richards, G., Vanden Berk, D., Reichard, T., et al. 2002, AJ, 124, 1 Richards et al. 2001, AJ, 121, 2308 Risaliti, G. & Elvis, M. 2005, ApJ, 629, L17 Risaliti, G., Elvis, M., Gilli, R., et al. 2003, ApJ, 587, L9 Rodr´ıguez-Ardila, A. & Mazzalay, X. 2006, MNRAS, L12 Schaye, J. 2001, ApJ, 562, L95 Schneider, D., Hall, P., Richards, G., et al. 2005, AJ, 130, 367 Schneider, D. P., Richards, G. T., Fan, X., et al. 2002, AJ, 123, 567 Scoville, N. Z., Evans, A. S., Thompson, R., et al. 2000, AJ, 119, 991 Shastri, P., Wilkes, B. J., Elvis, M., et al. 1993, ApJ, 410, 29 Smith, P. S., Schmidt, G. D., Hines, D. C., et al. 2003, ApJ, 593, 676 Spergel, D. N., Bean, R., Dor´e, O., et al. 2006, ApJ, submitted (astro ph/0603449) Steffen, A. T., Strateva, I., Brandt, W. N., et al. 2006, AJ, in press (astro-ph/0602407) Stern, D., Eisenhardt, P., Gorjian, V., et al. 2005, ApJ, 631, 163 Strateva, I., Brandt, N., Schneider, D., et al. 2005, AJ, 130, 387 Tozzi, P., Gilli, R., Mainieri, V., et al. 2006, A&A, in press (astro- ph/0602127) Ueda, Y., Akiyama, M., Ohta, K., et al. 2003, ApJ, 598, 886 Vanden Berk, D., Richards, G., Bauer, A., et al. 2001, AJ, 122, 549 Wild, V. & Hewett, P. C. 2005, MNRAS, 361, L30 Wilkes, B., Pounds, K., Schmidt, G., et al. 2005, ApJ, 634, 183 Wilkes, B., Schmidt, G., Cutri, R., et al. 2002, ApJ, 564, L65 Wills, B. J., Laor, A., Brotherton, M. S., et al. 1999, ApJ, 515, L53 Worsley, M., Fabian, A., Bauer, F., et al. 2005, MNRAS, 357, 1281 York, D., Adelman, J., Anderson, J., et al. 2000, AJ, 120, 1579 York, D., Khare, P., Vanden Berk, D., et al. 2006, MNRAS, 367, 945 X-RAYING RED QUASARS 9 Fig. 1. -- The g − i colors of the SDSS EDR quasar sample from which our targets were chosen. The contours show the distribution of g − i colors for all 3814 quasars from the EDR. Circles denote optically selected red quasars with ∆(g − i) > 0.2. The filled black circles are the 16 red quasars that met our other selection criteria: 0.6 ≤ z ≤ 2.2 (vertical dashed lines), i < 18, no broad absorption lines (BALs), radio-quiet, and low Galactic NH . (Of these, the 12 listed in Table 1 were awarded Chandra time.) Only ∼11.5% of non-BAL, i < 18 quasars in our redshift range had relative colors red enough to be considered as a candidate red quasar, but many of the reddest quasars are contaminated by host galaxy light (at z < 0.6) or are BAL quasars, either of which was grounds for exclusion from our sample. 10 HALL ET AL. Fig. 2. -- SDSS optical/UV spectra of our red quasar candidates. The gray curves in each panel show the Vanden Berk et al. (2001) composite quasar spectrum for reference. To further illustrate the difference between normal and red quasars, the dotted gray curves in the bottom right panel show power-law continua with αν = −0.5 (normal) and −1.0 (red). Quasars with negative values of ∆αox have their names shown in gray italics. The three of those quasars which have the hardest ΓHR values also have their redshifts shown in gray italics. X-RAYING RED QUASARS 11 Fig. 3. -- A plot of ∆(J − K) versus ∆(g − i) for SDSS quasars with 2MASS photometry. Both colors have been normalized to the median at each redshift and so most objects cluster around a value of zero in each normalized color, as shown by the contours (points show objects in regions of lower density than the lowest contour level). The ∆(g − i) distribution is shown along the bottom, along with a Gaussian overplotted as a dotted line to show a symmetric distribution. The red (positive) tail of the ∆(g − i) distribution is symptomatic of dust reddening (Richards et al. 2003), but the ∆(J − K) distribution is roughly symmetric and does not show such a tail. Furthermore, those quasars that are reddest in ∆(g − i) do not generally have anomalous ∆(J − K) colors. The inset shows J − K vs. ∆(g − i); note that the quasars reddest in ∆(g − i) do not generally meet a J − K > 2 selection criterion either. 12 HALL ET AL. Fig. 4. -- X-ray photon index, ΓHR, versus two different measures of the optical/UV color. Small points are from Gallagher et al. (2005) and span the range of "normal" quasars. Large symbols are our reddened quasar candidates; open symbols have ∆αox> 0. Triangles indicate quasars that have probable DLAs along their line of sight. Panel (a): Optical color as determined from the SDSS photometry using ∆(g − i). Most normal (unreddened) quasars have −0.2 <∆(g − i)< 0.2. Lines show X-ray absorption/optical reddening tracks for gas and dust with an SMC-like extinction curve at z = 1.5 but with the different labeled dust-to-gas ratios. The solid track has an SMC dust-to-gas ratio, the dashed track has a dust-to-gas ratio 0.1 times that of the SMC, and the dot-dashed track has a dust-to-gas ratio 0.01 times that of the SMC. The length of each track corresponds to E(B − V ) = 0.1, although the dot-dashed track only reaches E(B − V ) = 0.04 before going off the plot. Panel (b): Optical color as determined from the SDSS spectra by fitting a power-law to the continuum. The dashed line is the BCES(YX) fit to the full sample, while the dotted lines show the ±1σ uncertainties in the fit. X-RAYING RED QUASARS 13 Fig. 5. -- Luminosity-corrected optical-to-X-ray flux ratio, ∆αox, versus two different measures of the optical color. Small points with error bars are from Gallagher et al. (2005) and span the range of "normal" quasars. Large symbols are our reddened quasar candidates; open symbols have ∆αox> 0. Triangles indicate quasars that have probable DLAs along their line of sight. Small stars are from Strateva et al. (2005), representing SDSS quasars with sensitive ROSAT observations. The horizontal dashed lines show the ±1σ standard deviation of the Strateva et al. (2005) ∆αox distribution. Panel (a): Optical color as determined from the SDSS photometry using ∆(g − i). Lines show X-ray absorption/optical reddening tracks for gas and dust with an SMC-like extinction curve at z = 1.5 but with the different labeled dust-to-gas ratios. The solid track has an SMC dust-to-gas ratio, the dashed track has a dust-to-gas ratio 0.1 times that of the SMC, and the dot-dashed track has a dust-to-gas ratio 0.01 times that of the SMC. The length of each track corresponds to E(B − V ) = 0.1. Panel (b): Optical color as determined from the SDSS spectra by fitting a power-law to the continuum. 14 HALL ET AL. Fig. 6. -- The X-ray spectral index, ΓHR, (determined from the observed hardness ratio) vs. ∆αox, the optical/UV to X-ray index αox relative to the value expected for the quasar's luminosity. Small points with error bars are from Gallagher et al. (2005) and span the range of "normal" quasars. Large symbols are our reddened quasar candidates; open symbols have ∆αox> 0. Triangles indicate quasars that have probable DLAs along their line of sight. Dotted lines show the effects of dust-free neutral absorption at the redshifts indicated. The crosses on those lines mark the locations of column densities of NH = 2.5 × 1022 cm−2 at each redshift. The non-dotted lines show X-ray absorption/optical reddening tracks for gas with SMC-like dust at z = 1.5. The solid track has an SMC dust-to-gas ratio, the dashed track has a dust-to-gas ratio 0.1 times that of the SMC, and the dot-dashed track has a dust-to-gas ratio 0.01 times that of the SMC. The length of each track corresponds to E(B − V ) = 0.1, although the dot-dashed track only reaches E(B − V ) = 0.04 before going off the plot. X-RAYING RED QUASARS 15 Fig. 7. -- The previous Figure with the addition of Risaliti et al. (2003) objects (blue asterisks), using numbers taken from their Table 1 and assuming an expected αox=−1.73 for their luminosity range (their Figure 3). With that appropriate average αox instead of the value −1.55 they used, which is appropriate only for lower-luminosity objects like those found at z < 0.5 in the Bright Quasar Survey sample (Brandt et al. 2000), it is clear they have not found a new population of X-ray weak quasars. The blue cross marks the position of PHL 1811 (Choi et al. 2005), an interesting outlier in this parameter space. The magenta polygon outlines the location of the bulk of the BAL quasars from the sample of Gallagher et al. (2006). See §5 for further discussion of this Figure. Table 1 Target List Name (SDSS J) z i Mi ∆(u − r) ∆(g − i) ∆(r − z) αUV a NH b Spectrumc Commentsd 000230.71+004959.0 113345.62+005813.4 122652.01−001159.5 125140.32+000210.8 130211.04+000004.5 131058.13+010822.2 132323.78−002155.2 170817.85+615448.5 171419.23+611944.7 171535.96+622336.0 173551.92+535515.7 173836.16+583748.5 1.355 1.937 1.176 0.878 1.797 1.392 1.388 1.415 1.847 2.182 0.956 1.279 17.91 −26.54 17.86 −27.41 17.68 −26.43 17.43 −25.97 17.96 −27.16 17.80 −26.66 17.61 −26.90 17.64 −26.91 17.92 −27.27 17.96 −27.59 17.77 −25.84 17.51 −26.81 0.604±0.045 0.458±0.071 0.459±0.026 0.402±0.027 0.191±0.028 0.741±0.048 0.762±0.028 0.290±0.031 0.089±0.038 0.116±0.040 0.228±0.023 0.377±0.028 0.396±0.037 0.382±0.036 0.261±0.026 0.319±0.023 0.363±0.027 0.626±0.042 0.463±0.033 0.259±0.022 0.426±0.027 0.352±0.027 0.235±0.023 0.320±0.024 0.114±0.035 −0.97 0.050±0.065 −1.15 0.201±0.031 −0.72 0.185±0.030 −1.01 0.193±0.033 −0.86 0.444±0.029 −1.61 0.275±0.026 −1.12 0.108±0.032 −1.13 0.334±0.042 −0.97 0.296±0.036 −1.05 0.073±0.026 −0.25 0.185±0.028 −0.70 3.14 2.80 1.91 1.59 1.56 1.95 1.89 2.51 2.64 2.62 3.42 3.63 387-575-51791 282-408-51658 289-198-51990 292-559-51609 293-080-51994 295-325-51985 297-267-51959 351-550-51780 354-360-51792 352-499-51789 360-219-51816 366-038-52017 dust; DLA? dust/int?; DLA?, XRF dust/SED? dust; XRF SED; DLA?, XRF dust/int?; DLA, XRF int; DLA dust/SED? SED SED normal dust; XRF aThe spectral index for the ultraviolet continuum (where fν ∝ ναUV ) is measured by fitting the SDSS spectra. bThe values for NH (in units of 1020 cm−2) used in the X-ray simulations are from Galactic H I maps (Dickey & Lockman 1990). cThe SDSS spectrum of each target is identified by the code plate-fiber-MJD, giving its spectroscopic plate number, its fiber number within that plate, and the MJD on which the plate was observed. dThe first comment is our best explanation for the object's redness (see § 4.2): dust -- dust reddening at the quasar redshift; SED -- an intrinsically red SED; int -- dust reddening by an intervening absorber; normal -- apparently misclassified as red. Other comments: DLA?/DLA -- The SDSS spectrum shows weak/strong evidence for a damped Lyα absorber candidate; XRF -- X-ray faint source with αox more than 1σ from the αox-L2500 relation of Strateva et al. (2005), where σ is the RMS scatter in αox from Table 5 of Steffen et al. (2006). 1 6 H A L L E T A L . X-RAYING RED QUASARS 17 Table 2 Chandra Observing Log Name (SDSS J) 000230.71+004959.0 113345.62+005813.4 122652.01−001159.5 125140.32+000210.8 130211.04+000004.5 131058.13+010822.2 132323.78−002155.2 170817.85+615448.5 171419.23+611944.7 171535.96+632336.0 173551.92+535515.7 173836.16+583748.5 Obs. ID Date (MJD) Exposure Time (ks) Countsa Soft Hard Count Ratea (10−3 ct s−1) HRb 4861 4858 4865 4859 4862 4854 4855 4864 4856 4857 4863 4860 2003 Nov 09 (52952) 2004 Mar 29 (53093) 2004 Feb 12 (53047) 2004 Mar 08 (53072) 2004 May 12 (53150) 2004 Mar 26 (53090) 2004 May 09 (53147) 2003 Nov 16 (52959) 2003 Nov 16 (52959) 2004 Sep 29 (53277) 2003 Nov 09 (52952) 2003 Nov 16 (52959) 5.67 4.26 4.90 4.04 4.90 5.45 4.36 4.06 4.62 4.25 5.42 3.87 130+12.4 −11.4 < 3.8 87+10.4 −9.3 27+6.3 −5.2 25+6.1 −5.0 7+3.8 −2.6 102+11.1 −10.1 153+13.4 −12.4 32+6.7 −5.6 50+8.1 −7.0 188+14.7 −13.7 5+3.4 −2.2 20+5.6 −4.4 < 2.3 16+5.1 −4.0 12+4.6 −3.4 8+4.0 −2.8 5+3.4 −2.2 29+6.5 −5.4 46+7.8 −6.8 13+4.7 −3.6 12+4.6 −3.4 47+7.9 −6.8 < 5.3 26.63+2.35 −2.16 < 0.88 21.03+2.28 −2.07 9.89+1.83 −1.56 6.73+1.39 −1.17 2.39+0.86 −0.65 30.05+2.86 −2.62 49.02+3.73 −3.47 9.74+1.68 −1.45 14.81+2.11 −1.86 43.34+3.02 −2.83 1.81+0.98 −0.67 −0.73+0.07 −0.06 · · · −0.69+0.09 −0.07 −0.38+0.17 −0.16 −0.52+0.18 −0.16 −0.17+0.34 −0.32 −0.56+0.08 −0.07 −0.54+0.07 −0.06 −0.42+0.16 −0.14 −0.61+0.12 −0.10 −0.60+0.06 −0.05 < 0.03 aDetections for the full, soft, and hard bands are determined by wavdetect (as described in §3). Errors are 1σ Poisson errors (Gehrels 1986), except for non-detections where the limits are the 90% confidence limits from Bayesian statistics (Kraft, Burrows, & Nousek 1991). The count rate is for the full band, 0.5 -- 8.0 keV. bThe HR is defined as (h − s)/(h + s), where h and s are the counts in the hard (2.0 -- 8.0 keV) and soft (0.5 -- 2.0 keV) bands, respectively. Errors in the HR are propagated from the counting errors using the numerical method of Lyons (1991). Table 3 Derived Properties Name (SDSS J) a ΓHR log(FX)b log(f2keV)c log(L2keV)c log(f2500)c log(L2500)d αox 000230.71+004959.0 113345.62+005813.4 122652.01−001159.5 125140.32+000210.8 130211.04+000004.5 131058.13+010822.2 132323.78−002155.2 170817.85+615448.5 171419.23+611944.7 171535.96+632336.0 173551.92+535515.7 173836.16+583748.5 2.13+0.23 −0.23 · · · 2.01+0.24 −0.28 1.24+0.34 −0.32 1.52+0.44 −0.36 0.84+0.62 −0.59 1.63+0.18 −0.19 1.58+0.15 −0.16 1.32+0.31 −0.32 1.75+0.32 −0.29 1.73+0.14 −0.15 > 0.48 −12.868 ± 0.037 −30.577 ± 0.037 < −14.029 < −32.287 −12.942 ± 0.045 −30.729 ± 0.045 −13.059 ± 0.074 −31.199 ± 0.074 −13.309 ± 0.083 −31.233 ± 0.083 −13.552 ± 0.136 −31.913 ± 0.136 −12.688 ± 0.040 −30.599 ± 0.040 −12.468 ± 0.032 −30.400 ± 0.032 −13.087 ± 0.070 −31.131 ± 0.070 −13.032 ± 0.058 −30.791 ± 0.058 −12.557 ± 0.029 −30.483 ± 0.029 −13.581 ± 0.195 −32.176 ± 0.195 27.096 < 25.673 26.826 26.107 26.668 25.782 27.094 27.309 26.791 27.260 26.896 25.450 −26.691 −26.507 −26.514 −26.445 −26.536 −26.626 −26.514 −26.484 −26.609 −26.507 −26.533 −26.428 30.983 31.452 31.042 30.861 31.365 31.069 31.179 31.225 31.314 31.545 30.846 31.197 −1.49 ± 0.10 < −2.22 −1.62 ± 0.10 −1.83 ± 0.13 −1.81 ± 0.11 −2.03 ± 0.18 −1.57 ± 0.10 −1.50 ± 0.09 −1.74 ± 0.11 −1.65 ± 0.10 −1.51 ± 0.10 −2.20 ± 0.20 ∆αox e 0.11 < −0.56 0.01 −0.25 −0.16 −0.42 0.05 0.13 −0.10 0.02 0.07 −0.57 aΓHR is a coarse measure of the hardness of the X-ray spectrum determined by comparing the observed HR (see Table 2) to a simulated HR that takes into account temporal variations in the instrument response and the Galactic NH toward the target (see §3). bThe full-band X-ray flux, FX, has units of erg cm−2 s−1 and is calculated by integrating the power-law spectrum given by ΓHR and normalized by the full-band count rate from 0.5 -- 8.0 keV. The errors are derived from the 1σ errors in the full-band count rate. cX-ray and optical flux densities were measured at rest-frame 2 keV and 2500 A, respectively; units are erg cm−2 s−1 Hz−1. dThe rest-frame 2 keV and 2500 A luminosity densities (L2keV and L2500, respectively) have units of erg s−1 Hz−1. e∆αox is the difference between the observed αox and the expected αox calculated from Equation 6 of Strateva et al. (2005) using L2500. The general quasar population has an observed RMS scatter around ∆αox=0 of ±0.146 at 30 < log(L2500) < 31 (Table 5 of Steffen et al. 2006). 1 8 H A L L E T A L . X-RAYING RED QUASARS 19 Table 4 Results from X-ray Spectral Fittinga Name (SDSS J) Γ NH C-stat/ν (1022 cm−2) Total Counts 000230.71+004959.0 122652.01−001159.5 132323.78−002155.2 170817.85+615448.5 171535.96+632336.0 173551.92+535515.7 −0.37 −0.26 −0.40 2.21+0.43 2.15+0.29 2.07+0.42 1.81+0.30 1.67+0.41 1.73+0.21 −0.32 −0.30 −0.17 0.39+0.61 −0.39 < 0.14 1.30+0.87 0.46+0.54 −0.46 < 1.23 < 0.19 −0.78 214/512 210/512 230/512 280/51 183/51 270/51 151 103 131 197 62 232 aAll spectra were fit with a power-law model with intrinsic absorption at the quasar redshift. The errors quoted are 1σ (68% confidence; ∆C = 2.30 for two parameters of interest). The redshift and Galactic NH for each quasar are fixed to the appropriate values (see Table 1). 20 HALL ET AL. APPENDIX intrinsic scatter in ∆αox and ΓHR Between this paper and Gallagher et al. (2005), we have a sample of about two dozen objects with data of sufficient quality and uniformity to investigate the intrinsic dispersion in ∆αox and ΓHR among radio-quiet quasars at zem ∼ 1.5. Accounting for our measurement errors and excluding the three clearly anomalous objects with the most negative ∆αox, all of which we have classified as predominantly dust reddened, we find the intrinsic dispersion in ∆αox to be σ = 0.05 ± 0.02, assuming a Gaussian distribution. (The standard deviation in observed ∆αox in the sample of Strateva et al. (2005) is ±0.13, but they do not estimate the intrinsic dispersion.) This σ is equivalent to a dispersion of ±35% in the X-ray luminosity at a fixed UV luminosity. Scatter in the line-of-sight dust extinction will contribute to this dispersion, but its amplitude is consistent with it being entirely due to X-ray variability (Manners, Almaini, & Lawrence 2002). Our uncertainties already include a statistical correction for UV variability. As for ΓHR, excluding the two lower limits we find the observed dispersion in ΓHR to be σΓ = 0.26 ± 0.09, in good agreement with George et al. (2000) and Mateos et al. (2005). The dispersion in ΓHR around the correlation with αUV is much lower (σ = 0.12 ± 0.07). Nonetheless, it still may be larger than the dispersion in Γ attributable to time variability (σ ≃ 0.06 ± 0.02; Markowitz, Edelson, & Vaughan 2003). We agree with Risaliti & Elvis (2005) that the remaining dispersion in ΓHR is not attributable to differences in obscuration, because in Figure 4 even the bluest quasars show a scatter in ΓHR. Nor is it attributable to trends with redshift or luminosity, since ΓHR does not seem to correlate with either (Tozzi et al. 2006). In summary, we suggest that the observed dispersion in αox seen in quasars lacking UV absorption is entirely explained by the known correlation of αox with luminosity, the effects of flux variability, and a small contribution from dust extinction. On the other hand, only ∼60% of the observed dispersion in ΓHR is explained by the effects of spectral variability plus the observed correlation of ΓHR with αUV. The remaining dispersion may arise because ΓHR does correlate with numerous other optical and ultraviolet spectral properties (e.g., Laor et al. 1994; Wills et al. 1999). Those latter studies, among others, have suggested that a steep ΓHR corresponds to a large Eddington ratio. Thus, although the exact mechanisms which determine the observed values and trends of αox and ΓHR in quasars may be unclear, it appears that we have identified the predominant sources of scatter in those relationships.